Johannes A. Lercher

Find an error

Name: Johannes A. Lercher*
Organization: Technische Universit?t München , Germany
Department: Department Chemie
Title: Professor(PhD)

TOPICS

Co-reporter:Peter H. Hintermeier, Sebastian Eckstein, Donghai Mei, Mariefel V. Olarte, Donald M. Camaioni, Eszter Baráth, and Johannes A. Lercher
ACS Catalysis November 3, 2017 Volume 7(Issue 11) pp:7822-7822
Publication Date(Web):October 4, 2017
DOI:10.1021/acscatal.7b01582
Hydronium ions in the pores of zeolite H-ZSM5 show high catalytic activity in the elimination of water from cyclohexanol in aqueous phase. Substitution induces subtle changes in rates and reaction pathways, which are concluded to be related to steric effects. Exploring the reaction pathways of 2-, 3-, and 4-methylcyclohexanol (2-McyOH, 3-McyOH, and 4-McyOH), 2- and 4-ethylcyclohexanol (2-EcyOH and 4-EcyOH), 2-n-propylcyclohexanol (2-PcyOH), and cyclohexanol (CyOH) it is shown that the E2 character increases with closer positioning of the alkyl and hydroxyl groups. Thus, 4-McyOH dehydration proceeds via an E1-type elimination, while cis-2-McyOH preferentially reacts via an E2 pathway. The entropy of activation decreased with increasing alkyl chain length (ca. 20 J mol–1 K–1 per CH2 unit) for 2-substituted alcohols, which is concluded to result from constraints influencing the configurational entropy of the transition states.Keywords: Dehydration; Elimination reactions; Heterogeneous catalysis; Homogeneous catalysis; Zeolite;
Co-reporter:Stefan Schallmoser, Gary L. Haller, Maricruz Sanchez-Sanchez, and Johannes A. Lercher
Journal of the American Chemical Society June 28, 2017 Volume 139(Issue 25) pp:8646-8646
Publication Date(Web):June 6, 2017
DOI:10.1021/jacs.7b03690
The Brønsted acid sites of H-ZSM-5 and ferrierite reversibly adsborb linear pentenes via hydrogen bonding, rapidly isomerizing the double bond. On H-ZSM-5, dimerization of adsorbed pentenes occurs at a slower rate and leads to pentyl ester covalently bound to the surface. Pentene adsorbed on zeolites with narrower pores, such as ferrierite, remained stable in a hydrogen-bonded state even up to 423 K. Comparing the differential heat of adsorption of 2-pentene on silicalite and ferrierite allowed for the determination of the enthalpy difference between physically adsorbed pentene in ZSM-5 and the localized hydrogen-bonded π-complex at Brønsted acid sites, −36 kJ/mol. The activation energy (35 kJ/mol) for dimerization is almost identical to this enthalpy difference, suggesting that the rate-determining step is associated either with the mobilization of π-bonded 2-pentene or with the equally large activation barrier to form an alkoxy group via a carbenium-ion transition state. In a closed system, the dimerization rate is first order in the concentration of the π-complex that is both in equilibrium with the mobile pentene phase and in production of the carbenium ion that reacts with the mobile pentene. Overall, the alkoxy group is −41 ± 7 kJ/mol more stable than physisorbed pentene, establishing a series of energetically well-separated groups of reactive surface species.
Co-reporter:Zhenchao Zhao, Hui Shi, Chuan Wan, Mary Y. Hu, Yuanshuai Liu, Donghai Mei, Donald M. Camaioni, Jian Zhi Hu, and Johannes A. Lercher
Journal of the American Chemical Society July 12, 2017 Volume 139(Issue 27) pp:9178-9178
Publication Date(Web):June 19, 2017
DOI:10.1021/jacs.7b02153
The reaction mechanism of solid-acid-catalyzed phenol alkylation with cyclohexanol and cyclohexene in the apolar solvent decalin has been studied using in situ 13C MAS NMR spectroscopy. Phenol alkylation with cyclohexanol sets in only after a majority of cyclohexanol is dehydrated to cyclohexene. As phenol and cyclohexanol show similar adsorption strength, this strict reaction sequence is not caused by the limited access of phenol to cyclohexanol, but is due to the absence of a reactive electrophile as long as a significant fraction of cyclohexanol is present. 13C isotope labeling demonstrates that the reactive electrophile, the cyclohexyl carbenium ion, is directly formed in a protonation step when cyclohexene is the coreactant. In the presence of cyclohexanol, its protonated dimers at Brønsted acid sites hinder the adsorption of cyclohexene and the formation of a carbenium ion. Thus, it is demonstrated that protonated cyclohexanol dimers dehydrate without the formation of a carbenium ion, which would otherwise have contributed to the alkylation in the kinetically relevant step. Isotope scrambling shows that intramolecular rearrangement of cyclohexyl phenyl ether does not significantly contribute to alkylation at the aromatic ring.
Co-reporter:Manuel F. Wagenhofer, Eszter Baráth, Oliver Y. Gutiérrez, and Johannes A. Lercher
ACS Catalysis February 3, 2017 Volume 7(Issue 2) pp:1068-1068
Publication Date(Web):December 14, 2016
DOI:10.1021/acscatal.6b02753
The mechanism of the deoxygenation of fatty acids on transition-metal sulfides was determined on the basis of kinetic data obtained with fatty acids, their reaction intermediates (aldehyde and alcohol), and reactants of restricted reactivity (adamantanyl-substituted carboxylic acids). Deoxygenation on MoS2 proceeds exclusively via hydrogenolysis to aldehyde, followed by hydrogenation to the corresponding alcohol, consecutive dehydration to the olefin, and hydrogenation to the alkane. In contrast, the selectivity on Ni-MoS2 and on Ni3S2 substantially shifts toward carbon oxide elimination routes: i.e., direct production of Cn–1 olefins and alkanes. The carbon losses occur by decarbonylation of a ketene intermediate, which forms only on sites associated with Ni. The rate determining steps are the cleavage of the C–C bond and the removal of oxygen from the surface below and above, respectively, 2.5 MPa of H2. The different reaction pathways catalyzed by MoS2 and Ni-MoS2 are attributed to a preferred deprotonation of a surface acyl intermediate formed upon the adsorption of the fatty acid on Ni-MoS2. The shift in mechanism is concluded to originate from the higher basicity of sulfur induced by nickel.Keywords: decarbonylation; fatty acid; hydrodeoxygenation; reaction mechanism; transition metal sulfides;
Co-reporter:Ricardo Bermejo-Deval;Raimund M. H. Walter;Oliver Y. Gutiérrez
Catalysis Science & Technology (2011-Present) 2017 vol. 7(Issue 19) pp:4437-4443
Publication Date(Web):2017/10/02
DOI:10.1039/C7CY01255A
The electronegativity effect of the alkali cations on the formation of methanethiol by reaction of methanol and H2S was studied with K+, Rb+, and Cs+ supported on γ-Al2O3. Catalyst activation in H2S led to the formation of sulfur oxyanions, including SO3−2, S2O3−2, SO4−2 and S2O6−2. The surface modified aluminas have strong basic properties. Accessible alkali cations on γ-Al2O3 are the active sites catalyzing the SN2 nucleophilic substitution. These sites stabilize two Lewis acid–base pairs, formed upon H2S and methanol dissociation, in proximity. Decreasing electronegativity of the alkali enhances the concentration of basic sites and their adsorption capacity. In consequence, the rates of methanethiol formation increased in the sequence K/γ-Al2O3 < Rb/γ-Al2O3 < Cs/γ-Al2O3. The activation energies also decreased in parallel with the decreasing electronegativity of the alkali, which is attributed to the increasing electron density at the SH− groups that facilitates the nucleophilic attack.
Co-reporter:Eva Schachtl, Jong Suk Yoo, Oliver Y. Gutiérrez, Felix Studt, Johannes A. Lercher
Journal of Catalysis 2017 Volume 352(Volume 352) pp:
Publication Date(Web):1 August 2017
DOI:10.1016/j.jcat.2017.05.003
•Kinetic study of the hydrogenation of polyaromatic compounds on (Ni-)MoS2/Al2O3.•Ni decorating the MoS2 phase increases its dispersion and concentration of active H.•Reduced activation barriers accelerate H addition to the aromatic on promoted edges.•Rate determining step (L-H mechanism) shifts from the first to the second H-pair addition.•Thus, hydrogenation is faster and deeper on Ni-promoted sites than on unpromoted sites.The reaction network and elementary steps of the hydrogenation of phenanthrene are explored on parent and Ni-promoted MoS2/γ-Al2O3. Two pathways were identified, i.e., Path 1: Phenanthrene ⇌ 9,10-dihydrophenanthrene (DiHPhe) → 1,2,3,4,4a,9,10,10a-octahydro-phenanthrene (asymOHPhe), and Path 2: Phenanthrene → 1,2,3,4-tetrahydrophenanthrene (TetHPhe) → 1,2,3,4,5,6,7,8-octahydrophenanthrene. The steps TetHPhe → asymOHPhe (hydrogenation), and DiHPhe → TetHPhe (hydrogenation-isomerization) become notable at phenanthrene conversions above 20%. The reaction preferentially proceeds via Path 1 (90% selectivity) on MoS2/Al2O3. Ni promotion (Ni/(Ni + Mo) molar ratio of 0.3 at the edges on MoS2) increases the hydrogenation activity per active edge twofold and leads to 50% selectivity to both pathways. The reaction orders in H2 vary from ∼0.8 on MoS2/Al2O3 to ∼1.2 on Ni-MoS2/Al2O3, whereas the reaction orders in phenanthrene (∼0.6) hardly depend on Ni promotion. The reaction orders in H2S are zero on MoS2/Al2O3 and slightly negative on Ni-MoS2/Al2O3. DFT calculations indicate that phenanthrene is preferentially adsorbed parallel to the basal planes, while H is located at the edges perpendicular to the basal planes. Theory also suggests that Ni atoms, incorporated preferentially on the S-edges, increase the stability of hydrogenated intermediates. Hydrogenation of phenanthrene proceeds through quasi-equilibrated adsorption of the reactants followed by consecutive addition of hydrogen pairs to the adsorbed hydrocarbon. The rate determining steps for the formation of DiHPhe and TetHPhe are the addition of the first and second hydrogen pair, respectively. The concentration of SH groups (activated H at the edges) increases with Ni promotion linearly correlating the rates of Path 1 and Path 2, albeit with different functions. The enhancing effect of Ni on Path 2 is attributed to accelerated hydrogen addition to adsorbed hydrocarbons without important changes in their coverages.Download high-res image (131KB)Download full-size image
Co-reporter:Sylvia Albersberger, Jennifer Hein, Moritz W. Schreiber, Santiago Guerra, Jinyi Han, Oliver Y. Gutiérrez, Johannes A. Lercher
Catalysis Today 2017 Volume 297(Volume 297) pp:
Publication Date(Web):15 November 2017
DOI:10.1016/j.cattod.2017.05.083
•Unsupported Ni-Mo-W sulfides obtained from precursors synthesized via different routes.•Hydrodenitrogenation (HDN) and hydrodesulfurization (HDS) of model compounds.•Abundant exposed active sites indicated by high specific surface area and NO adsorption.•High W content needed for high HDN and HDS activity in the presence of quinoline.•Short Mo(W)S2 slabs and homogenous Ni distribution enhance HDS and HDN activity.The catalytic properties of unsupported Ni-Mo-W sulfides (composites of Ni-Mo(W)S2 mixed sulfides and Ni3S2) obtained from precursors synthesized via co-precipitation, hydrothermal, and thiosalt decomposition were explored for hydrodenitrogenation (HDN) of o-propylaniline or quinoline in presence and absence of dibenzothiophene undergoing hydrodesulfurization (HDS). The conversion rates varied with the reacting substrate and were related to specific catalyst properties such as morphology, texture, surface and composition. For the HDN of o-propylaniline and the HDS of dibenzothiophene in presence of o-propylaniline, high concentrations of coordinatively unsaturated cationic sites (as characterized by NO adsorption) and the specific surface areas determined the rates of reaction. For the HDN of quinoline and the HDS of dibenzothiophene in the presence of quinoline, the high hydrogenation activity of tungsten sulfide and length of the slabs was found to be more important than in the cases with o-propylaniline. Overall, the activity of unsupported catalysts relates to the size of sulfide slabs provided that Ni is present at the perimeter of the slabs.Download high-res image (302KB)Download full-size image
Co-reporter:Sebastian Müller, Yue Liu, Felix M. Kirchberger, Markus Tonigold, Maricruz Sanchez-Sanchez, and Johannes A. Lercher
Journal of the American Chemical Society 2016 Volume 138(Issue 49) pp:15994-16003
Publication Date(Web):November 16, 2016
DOI:10.1021/jacs.6b09605
Hydrogen transfer is the major route in catalytic conversion of methanol to olefins (MTO) for the formation of nonolefinic byproducts, including alkanes and aromatics. Two separate, noninterlinked hydrogen transfer pathways have been identified. In the absence of methanol, hydrogen transfer occurs between olefins and naphthenes via protonation of the olefin and the transfer of the hydride to the carbenium ion. A hitherto unidentified hydride transfer pathway involving Lewis and Brønsted acid sites dominates as long as methanol is present in the reacting mixture, leading to aromatics and alkanes. Experiments with purely Lewis acidic ZSM-5 showed that methanol and propene react on Lewis acid sites to HCHO and propane. In turn, HCHO reacts with olefins stepwise to aromatic molecules on Brønsted acid sites. The aromatic molecules formed at Brønsted acid sites have a high tendency to convert to irreversibly adsorbed carbonaceous deposits and are responsible for the critical deactivation in the methanol to olefin process.
Co-reporter:Nirala Singh, Yang Song, Oliver Y. Gutiérrez, Donald M. Camaioni, Charles T. Campbell, and Johannes A. Lercher
ACS Catalysis 2016 Volume 6(Issue 11) pp:7466
Publication Date(Web):October 4, 2016
DOI:10.1021/acscatal.6b02296
Both electrocatalytic hydrogenation (ECH) and thermal catalytic hydrogenation (TCH) of phenol by Pt and Rh show a rollover in rate with increasing temperature without changing the principal reaction pathways. The negative effect of temperature for aqueous-phase phenol TCH and ECH on Pt and Rh is deduced to be from dehydrogenated phenol adsorbates, which block active sites. ECH and TCH rates increase similarly with increasing hydrogen chemical potential whether induced by applied potential or H2 pressure, both via increasing H coverage, and indirectly by removing site blockers, a strong effect at high temperature. This enables unprecedented phenol TCH rates at 60–100 °C.Keywords: electrocatalysis; hydrogenation; phenol; platinum; poisoning; rhodium
Co-reporter:Navneet K. Gupta, Bo Peng, Gary L. Haller, Erika E. Ember, and Johannes A. Lercher
ACS Catalysis 2016 Volume 6(Issue 9) pp:5843
Publication Date(Web):July 11, 2016
DOI:10.1021/acscatal.6b01424
The carbon-catalyzed reaction of Cl2 and CO constitutes the most important industrial route to phosgene. Although defects in carbon lead to surface chemical reactions, direct polarization of C-heteroatom bonds induces a more successful Cl2 catalytic activation, the rate-determining step in the overall catalytic cycle. The interplay between the electron-donating and -withdrawing ability of the incorporated nitrogen substituents on the formation and stabilization of active sites was examined by X-ray photoelectron and Raman spectroscopy. Mechanistic studies indicate that the polarized Cl2 induced by the direct interaction of Cl2 with a strongly electron-deficient carbon site in close proximity to a nitrogen substituent is essential for phosgene production. Nitrogen substitution into ordered carbon materials led to very active and stable carbon catalysts for COCl2 synthesis.Keywords: Cl2 activation; electrophilic and nucleophilic carbon sites; in situ Raman spectroscopy; N-containing carbon catalyst; reaction mechanisms; transient COCl2 synthesis
Co-reporter:Wenji Song, Yuanshuai Liu, Eszter Baráth, Lucy L. Wang, Chen Zhao, Donghai Mei, and Johannes A. Lercher
ACS Catalysis 2016 Volume 6(Issue 2) pp:878
Publication Date(Web):December 23, 2015
DOI:10.1021/acscatal.5b01217
Liquid-phase dehydration of 1-octadecanol, which is intermediately formed during the hydrodeoxygenation of microalgae oil on zeolite H-BEA, has been studied, combining experiment and theory. Both the OH group and the alkyl chain of 1-octadecanol interact with zeolite Brønsted acid sites, inducing inefficient utilization in the presence of high acid-site concentrations. The parallel intramolecular and intermolecular dehydration pathways, leading to octadecene and dioctadecyl ether, have different activation energies and pass through different reaction intermediates. The formation of surface alkoxides is the rate-limiting step in the intramolecular dehydration, whereas the intermolecular dehydration proceeds via a bulky dimer intermediate, occurring preferentially at the pore mouth or outer surface of zeolite crystallites. Despite the main contribution of Brønsted acid sites toward both dehydration pathways, Lewis acid sites are also active to form dioctadecyl ether.Keywords: 1-octadecanol; Brønsted acid site; dehydration; density functional theory; H-BEA zeolite; Lewis acid site
Co-reporter:S. Grundner, W. Luo, M. Sanchez-Sanchez and J. A. Lercher  
Chemical Communications 2016 vol. 52(Issue 12) pp:2553-2556
Publication Date(Web):24 Dec 2015
DOI:10.1039/C5CC08371K
Cu-Exchanged zeolites are known as active materials for methane oxidation to methanol. However, understanding of the formation of Cu active species during synthesis, dehydration and activation is fragmented and rudimentary. We show here how a synthesis protocol guided by insight in the ion exchange elementary steps leads to highly uniform Cu species in mordenite (MOR).
Co-reporter:Monica A. C. Markovits;Andreas Jentys;Moniek Tromp
Topics in Catalysis 2016 Volume 59( Issue 17-18) pp:1554-1563
Publication Date(Web):2016 October
DOI:10.1007/s11244-016-0676-x
A series of Cu/ZSM-5 materials were synthesized and tested for the selective oxidation of methane to methanol reaction in a three stage reaction. The efficiency of the catalysts is related to the ability of the zeolite framework to stabilize multinuclear Cu-oxo species, namely dicopper and tricopper oxo clusters. Spectroscopy characterization by EXAFS showed that the exchange with moderate Cu loadings led to preferential formation of trinuclear Cu complexes [Cu3(μ–O)3]2+ in HZSM-5. The concentration of Al pairs in ZSM-5 is found to limit the maximum concentration of multinuclear Cu-oxo species that can be formed. Above such maximum, inactive Cu species including Cu oxide nanoparticles are formed. Conversely, it was found that at low loadings the Cu speciation in Cu/ZSM-5 occurs as a mixture of Cu monomers and dimers. Furthermore, it was found that not only the structure of Cu-oxo clusters is relevant for the activation of methane, but also the local environment in which the cluster is embedded. Comparison of methane to Cu stoichiometries achieved for Cu/ZSM-5 and Cu/MOR systems containing the same type of active [Cu3(μ–O)3]2+ cluster shows that approximately 50 % of these clusters are inactive on ZSM-5. While MOR stabilizes the trinuclear clusters in highly constrained 8-MR side pockets, the possibility of ZSM-5 to stabilize part of these clusters in less constrained local environments might be the reason for a lower activity in methane oxidation.
Co-reporter:Yuchun Zhi; Hui Shi; Linyu Mu; Yue Liu; Donghai Mei; Donald M. Camaioni
Journal of the American Chemical Society 2015 Volume 137(Issue 50) pp:15781-15794
Publication Date(Web):November 11, 2015
DOI:10.1021/jacs.5b09107
The Brønsted acid-catalyzed gas-phase dehydration of 1-propanol (0.075–4 kPa) was studied on zeolite H-MFI (Si/Al = 26, containing minimal amounts of extra framework Al moieties) in the absence and presence of co-fed water (0–2.5 kPa) at 413–443 K. It is shown that propene can be formed from monomeric and dimeric adsorbed 1-propanol. The stronger adsorption of 1-propanol relative to water indicates that the reduced dehydration rates in the presence of water are not a consequence of the competitive adsorption between 1-propanol and water. Instead, the deleterious effect is related to the different extents of stabilization of adsorbed intermediates and the relevant elimination/substitution transition states by water. Water stabilizes the adsorbed 1-propanol monomer significantly more than the elimination transition state, leading to a higher activation barrier and a greater entropy gain for the rate-limiting step, which eventually leads to propene. In a similar manner, an excess of 1-propanol stabilizes the adsorbed state of 1-propanol more than the elimination transition state. In comparison with the monomer-mediated pathway, adsorbed dimer and the relevant transition states for propene and ether formation are similarly, while less effectively, stabilized by intrazeolite water molecules.
Co-reporter:Aleksei Vjunov, John L. Fulton, Donald M. Camaioni, Jian Z. Hu, Sarah D. Burton, Ilke Arslan, and Johannes A. Lercher
Chemistry of Materials 2015 Volume 27(Issue 9) pp:3533
Publication Date(Web):April 8, 2015
DOI:10.1021/acs.chemmater.5b01238
In this work, Al K-edge extended X-ray absorption fine structure and 27Al MAS NMR spectroscopies in combination with DFT calculations are used to determine both qualitative and quantitative structural changes of two well-characterized samples with the BEA structure. The effects of various properties, including Al concentration, Al distribution, particle size, and structural defects, on zeolite stability are explored. As the samples are degraded by treatment in hot liquid water, the local structure about the Al T-site remains mostly intact, including the Al–O–Si angles and bond distances, while the crystalline structure as measured by XRD and STEM is disrupted. The combined data suggests the crystallinity decreases via selective hydrolysis of the T1- and T2-sites that form the 4-member rings of the zeolite framework. The hydrolysis eventually leads to the dissolution of the T-sites followed by reprecipitation on the particle surface resulting in amorphization of the sample.
Co-reporter:Stanislav Kasakov, Hui Shi, Donald M. Camaioni, Chen Zhao, Eszter Baráth, Andreas Jentys and Johannes A. Lercher  
Green Chemistry 2015 vol. 17(Issue 11) pp:5079-5090
Publication Date(Web):05 Oct 2015
DOI:10.1039/C5GC02160J
Mechanistic aspects of deconstruction and hydrodeoxygenation of organosolv lignin using zeolite (HZSM-5 and HBEA) and SiO2 supported Ni catalysts are reported. Lignin was deconstructed and converted to substituted alicyclic and aromatic hydrocarbons with 5 to 14 carbon atoms. Full conversion with total yield of 70 ± 5 wt% hydrocarbons was achieved at 593 K and 20 bar H2. The organosolv lignin used consists of seven to eight monolignol subunits and has an average molecular weight of ca. 1200 g mol−1. The monolignols were mainly guaiacyl, syringyl and phenylcoumaran, randomly interconnected through β-O-4, 4-O-5, β-1, 5-5′ and β–β ether bonds. In situ IR spectroscopy was used to follow the changes in lignin constituents during reaction. The reductive catalytic deconstruction of organosolv lignin starts with the hydrogenolysis of aryl alkyl ether bonds, followed by hydrogenation of the aromatic rings to cyclic alcohols. Oxygen is removed from the alcohols via dehydration on Brønsted acid sites to cyclic alkenes that are further hydrogenated.
Co-reporter:Wenji Song, Yuanshuai Liu, Eszter Baráth, Chen Zhao and Johannes A. Lercher  
Green Chemistry 2015 vol. 17(Issue 2) pp:1204-1218
Publication Date(Web):10 Nov 2014
DOI:10.1039/C4GC01798F
The cleavage of C–O bonds in phenol, catechol, and guaiacol has been explored with mono- and dual-functional catalysts containing Ni and/or HZSM-5 in the aqueous phase. The aromatic ring of phenol is hydrogenated in the first step, and the C–O bond of the resulting cyclohexanol is dehydrated in sequence. The initial turnover frequency (TOF) of phenol hydrodeoxygenation increases in parallel with the acid site concentration irrespective of the concentration of the accessible surface Ni atoms. For catechol and guaiacol conversion, Ni catalyzes the hydrogenolysis of the C–O bonds in addition to arene hydrogenation. For catechol, the hydrogenation of the aromatic ring and the hydrogenolysis of the phenolic –OH group occur in parallel with a ratio of 8:1. The saturated cyclohexane-1,2-diol can be further dehydrated over HZSM-5 or hydrogenolyzed on Ni to complete hydrodeoxygenation. Guaiacol undergoes primarily hydrogenolysis (75%) to phenol via demethoxylation, and the hydrogenation route accounts for only 25%. This is attributed to the steric effects arising from the adjacent sp3 hybrid O–CH3 group. 2-Methoxycyclohexanol (from guaiacol hydrogenation) reacts further either via hydrogenolysis by Ni to cyclohexanol or via acid catalyzed demethoxylation and rearrangement steps followed by the subsequent hydrogenation of the intermediately formed olefins. On Ni/HZSM-5, the hydrodeoxygenation activities are much higher for the phenolic monomers than for their respective saturated analogues, pointing to the importance of sp2 orbitals. The presence of proximal acid sites increases the activities of Ni in the presence of H2 by a synergistic action.
Co-reporter:Jiechen Kong, Mingyuan He, Johannes A. Lercher and Chen Zhao  
Chemical Communications 2015 vol. 51(Issue 99) pp:17580-17583
Publication Date(Web):28 Sep 2015
DOI:10.1039/C5CC06828B
The utilization of lignin as a fuel precursor has attracted attention, and a novel and facile process has been developed for one-pot conversion of lignin into cycloalkanes and alkanes with Ni catalysts under moderate conditions. This cascade hydrodeoxygenation approach may open the route to a new promising technique for direct liquefaction of lignin to hydrocarbons.
Co-reporter:Stefanie Simson
The Journal of Physical Chemistry C 2015 Volume 119(Issue 5) pp:2471-2482
Publication Date(Web):January 14, 2015
DOI:10.1021/jp510089s
SiO2-supported Pd–Au catalysts with Pd/Au molar ratios varying from 0.8 to 7.0 were used as catalysts for vinyl acetate synthesis under industrial conditions. Continued operation of the bimetallic catalysts at 150 °C led to the formation of the Pd1Au1 phase in the particles, with the remaining Pd atoms forming Pd nanoparticles by leaching of Pd as acetate. The presence of these phases was monitored by X-ray absorption spectroscopy (XAS) of the used catalysts. Temperature-resolved in situ XRD of the reduced samples in an inert atmosphere confirmed the phase separation into a Pd-rich phase and a Au-rich phase above 160 °C. CO adsorption and XRD of the catalysts used at 180 °C showed that phase separation also took place during vinyl acetate synthesis. The pronounced temperature dependence of the morphology and surface composition of the bimetallic Pd–Au catalysts determines the selectivity; the activity; and, in particular, the stability during vinyl acetate synthesis.
Co-reporter:GiovanniMaria Piccini
The Journal of Physical Chemistry C 2015 Volume 119(Issue 11) pp:6128-6137
Publication Date(Web):March 10, 2015
DOI:10.1021/acs.jpcc.5b01739
Heats of adsorption of methane, ethane, and propane in H-chabazite (Si/Al = 14.4) have been measured and entropies have been derived from adsorption isotherms. For these systems quantum chemical ab initio calculations of Gibbs free energies have been performed. The deviations from the experimental values for methane, ethane, and propane are below 3 kJ/mol for the enthalpy, and the Gibbs free energy. A hybrid high-level (MP2/CBS): low-level (DFT+dispersion) method is used to determine adsorption structures and energies. Vibrational entropies and thermal enthalpy contributions are obtained from vibrational partition functions for the DFT+dispersion potential energy surface. Anharmonic corrections have been evaluated for each normal mode separately. One-dimensional Schrödinger equations are solved for potentials obtained by (curvilinear) distortions of the normal modes using a representation in internal coordinates.
Co-reporter:Johannes A. Lercher
ACS Central Science 2015 Volume 1(Issue 7) pp:350
Publication Date(Web):October 7, 2015
DOI:10.1021/acscentsci.5b00311
Co-reporter:Eva Schachtl; Elena Kondratieva; Oliver Y. Gutiérrez
The Journal of Physical Chemistry Letters 2015 Volume 6(Issue 15) pp:2929-2932
Publication Date(Web):July 9, 2015
DOI:10.1021/acs.jpclett.5b01217
The activation of H2 and H2S on (Ni)MoS2/Al2O3 leads to the formation of SH groups with acid character able to protonate 2,6-dimethylpyridine. The variation in concentrations of SH groups induced by H2 and H2S adsorption shows that both molecules dissociate on coordinatively unsaturated cations and neighboring S2–. In the studied materials, one sulfur vacancy and four SH groups per 10 metal atoms exist at the active edges of MoS2 under the conditions studied. H2–D2 exchange studies show that Ni increases the concentration of active surface hydrogen by up to 30% at the optimum Ni loading, by increasing the concentration of H2 and H2S chemisorption sites.
Co-reporter:Stanislav Kasakov;Dr. Chen Zhao;Dr. Eszter Baráth;Zizwe A. Chase;Dr. John L. Fulton;Dr. Donald M. Camaioni;Aleksei Vjunov;Dr. Hui Shi;Dr. Johannes A. Lercher
Chemistry - A European Journal 2015 Volume 21( Issue 4) pp:1567-1577
Publication Date(Web):
DOI:10.1002/chem.201405242

Abstract

Sulfonated carbons were explored as functionalized supports for Ni nanoparticles to hydrodeoxygenate (HDO) phenol. Both hexadecane and water were used as solvents. The dual-functional Ni catalysts supported on sulfonated carbon (Ni/C-SO3H) showed high rates for phenol hydrodeoxygenation in liquid hexadecane, but not in water. Glucose and cellulose were precursors to the carbon supports. Changes in the carbons resulting from sulfonation of the carbons resulted in variations of carbon sheet structures, morphologies and the surface concentrations of acid sites. While the C-SO3H supports were active for cyclohexanol dehydration in hexadecane and water, Ni/C-SO3H only catalysed the reduction of phenol to cyclohexanol in water. The state of 3–5 nm grafted Ni particles was analysed by in situ X-ray absorption spectroscopy. The results show that the metallic Ni was rapidly formed in situ without detectable leaching to the aqueous phase, suggesting that just the acid functions on Ni/C-SO3H are inhibited in the presence of water. Using in situ IR spectroscopy, it was shown that even in hexadecane, phenol HDO is limited by the dehydration step. Thus, phenol HDO catalysis was further improved by physically admixing C-SO3H with the Ni/C-SO3H catalyst to balance the two catalytic functions. The minimum addition of 7 wt % C-SO3H to the most active of the Ni/C-SO3H catalysts enabled nearly quantitative conversion of phenol and the highest selectivity (90 %) towards cyclohexane in 6 h, at temperatures as low as 473 K, suggesting that the proximity to Ni limits the acid properties of the support.

Co-reporter:Sebastian Foraita;Dr. John L. Fulton;Zizwe A. Chase;Aleksei Vjunov;Pinghong Xu;Dr. Eszter Baráth;Dr. Donald M. Camaioni;Dr. Chen Zhao;Dr. Johannes A. Lercher
Chemistry - A European Journal 2015 Volume 21( Issue 6) pp:2423-2434
Publication Date(Web):
DOI:10.1002/chem.201405312

Abstract

The role of the specific physicochemical properties of ZrO2 phases on Ni/ZrO2 has been explored with respect to the reduction of stearic acid. Conversion on pure m-ZrO2 is 1.3 times more active than on t-ZrO2, whereas Ni/m-ZrO2 is three times more active than Ni/t-ZrO2. Although the hydrodeoxygenation of stearic acid can be catalyzed solely by Ni, the synergistic interaction between Ni and the ZrO2 support causes the variations in the reaction rates. Adsorption of the carboxylic acid group on an oxygen vacancy of ZrO2 and the abstraction of the α-hydrogen atom with the elimination of the oxygen atom to produce a ketene is the key to enhance the overall rate. The hydrogenated intermediate 1-octadecanol is in turn decarbonylated to heptadecane with identical rates on all catalysts. Decarbonylation of 1-octadecanol is concluded to be limited by the competitive adsorption of reactants and intermediate. The substantially higher adsorption of propionic acid demonstrated by IR spectroscopy and the higher reactivity to O2 exchange reactions with the more active catalyst indicate that the higher concentration of active oxygen defects on m-ZrO2 compared to t-ZrO2 causes the higher activity of Ni/m-ZrO2.

Co-reporter:Christian A. Gärtner ; André C. van Veen
Journal of the American Chemical Society 2014 Volume 136(Issue 36) pp:12691-12701
Publication Date(Web):August 13, 2014
DOI:10.1021/ja505411s
Ethane is oxidatively dehydrogenated with a selectivity up to 95% on catalysts comprising a mixed molten alkali chloride supported on a mildly redox-active Dy2O3-doped MgO. The reactive oxyanionic OCl– species acting as active sites are catalytically formed by oxidation of Cl– at the MgO surface. Under reaction conditions this site is regenerated by O2, dissolving first in the alkali chloride melt, and in the second step dissociating and replenishing the oxygen vacancies on MgO. The oxyanion reactively dehydrogenates ethane at the melt–gas phase interface with nearly ideal selectivity. Thus, the reaction is concluded to proceed via two coupled steps following a Mars-van-Krevelen-mechanism at the solid–liquid and gas–liquid interface. The dissociation of O2 and/or the oxidation of Cl– at the melt–solid interface is concluded to have the lowest forward rate constants. The compositions of the oxide core and the molten chloride shell control the catalytic activity via the redox potential of the metal oxide and of the OCl–. Traces of water may be present in the molten chloride under reaction conditions, but the specific impact of this water is not obvious at present. The spatial separation of oxygen and ethane activation sites and the dynamic rearrangement of the surface anions and cations, preventing the exposure of coordinatively unsaturated cations, are concluded to be the origin of the surprisingly high olefin selectivity.
Co-reporter:Florian Schüßler, Stefan Schallmoser, Hui Shi, Gary L. Haller, Erika Ember, and Johannes A. Lercher
ACS Catalysis 2014 Volume 4(Issue 6) pp:1743
Publication Date(Web):April 15, 2014
DOI:10.1021/cs500200k
La3+ cations exchanged into ultrastable zeolite Y and zeolite X promote catalytic isomerization, cracking, and alkylation of alkanes. La3+ cations stabilize the zeolite lattices and, more importantly, polarize alkane C–H bonds to enhance the rates of all three reactions. This unique activity leads to stable cracking and isomerization of reactive alkanes, with polarizable C–H bonds with adjacent tertiary or quaternary carbon atoms below 370 K. The presence of La3+ cations also enhances the zeolite catalyzed hydride transfer rate for isobutane alkylation with 2-butene leading to high catalyst stability. Solid state MAS NMR shows that the strongest positive effects are associated with nonhydroxylated La3+ cations accessible to the reacting molecules in supercages of the zeolite. The high activity is the result of a cooperative polarization of C–H bonds of alkanes by La3+ cations and the presence of stable and strong Brønsted acid sites.Keywords: alkylation; cracking; C−H bond polarization; hydride transfer; La3+ exchange
Co-reporter:Oliver Y. Gutiérrez, Srujan Singh, Eva Schachtl, Jeongnam Kim, Elena Kondratieva, Jennifer Hein, and Johannes A. Lercher
ACS Catalysis 2014 Volume 4(Issue 5) pp:1487
Publication Date(Web):April 11, 2014
DOI:10.1021/cs500034d
MoS2 and Ni-promoted MoS2 catalysts supported on γ-Al2O3, siliceous SBA-15, and Zr- and Ti-modified SBA-15 were explored for the simultaneous hydrodesulfurization (HDS) of dibenzothiophene (DBT) and hydrodenitrogenation (HDN) of o-propylaniline (OPA). In all cases, OPA reacted preferentially via initial hydrogenation, and DBT was converted through direct sulfur removal. HDN and HDS activities of MoS2 catalysts are determined by the dispersion of the sulfide phase. Ni promotion increased its dispersion and activity for DBT HDS and also increased the rate of HDN via enhancing the rate of hydrogenation. On nonpromoted MoS2 catalysts, HDS was strongly inhibited by NH3, and the addition of Ni dramatically reduced this inhibiting effect. The conclusion is that HDS is proportional to the concentration of Mo and Ni on the edges of sulfide particles. In contrast, the direct hydrodenitrogenation of OPA occurs only on accessible Mo cations and, hence, decreases with increasing Ni substitution. The nature of the support influences the dispersion of the nonpromoted catalysts as well as the decoration degree of Ni on the edges of the Ni–Mo–S phase. Furthermore, the acidity of the support influences the acidity of the supported sulfide phase, which may play an important role in HDN.Keywords: HDN; HDS; MoS2; Ni promotion; SBA-15; support effect
Co-reporter:Hui Shi ; Oliver Y. Gutiérrez ; Anmin Zheng ; Gary L. Haller
The Journal of Physical Chemistry C 2014 Volume 118(Issue 36) pp:20948-20958
Publication Date(Web):August 18, 2014
DOI:10.1021/jp505483e
H/D isotope effects on methylcyclohexane hydrogenolysis over Ir/Al2O3 catalysts were examined by combining measured rates with theoretical estimates provided by partition function based analyses. Normal H/D isotope effects (rH/rD > 1) were observed for endocyclic and exocyclic C–C bond hydrogenolysis. Hydrogenolysis is concluded to occur via stepwise dehydrogenation followed by cleavage of the C–C bond and subsequent hydrogenation of the cleaved entities. The so-called “multiplet” mechanism (i.e., the C–C bond of a flat-lying physisorbed cyclic molecule is cleaved upon the attack of a coadsorbed H atom) is unequivocally excluded. For ring-opening, either C–C bond cleavage or C–H(D) bond reformation may be rate-determining, due to their indistinguishable isotope effects under the studied conditions. C–H(D) bond dissociation does not control the rate of C–C bond hydrogenolysis. For the exocyclic cleavage of the methyl group, a higher degree of unsaturation of the surface intermediate and the potential impact of mobile H atoms on large Ir particles are noted.
Co-reporter:Dipl.-Chem. Jennifer Hein;Dr. Ana Hrabar;Dr. Andreas Jentys;Dr. Oliver Y. Gutiérrez;Dr. Johannes A. Lercher
ChemCatChem 2014 Volume 6( Issue 2) pp:
Publication Date(Web):
DOI:10.1002/cctc.201490008
Co-reporter:Miroslaw Derewinski ; Priit Sarv ; Xianyong Sun ; Sebastian Müller ; Andre C. van Veen
The Journal of Physical Chemistry C 2014 Volume 118(Issue 12) pp:6122-6131
Publication Date(Web):February 10, 2014
DOI:10.1021/jp4053677
Quantitative NMR spectroscopic studies (27Al, 31P, and 1H magic-angle spinning NMR) were combined with temperature-programmed desorption (TPD) to investigate the impacts of sequential treatment procedures, i.e., calcination and hot water washing, of phosphorus-modified ZSM-5 samples on their local structural changes for aluminum and phosphorus as well as on the acidities. Introduction of phosphorus resulted mainly in the transformation of tetrahedrally coordinated framework aluminum Altetra-frame to distorted aluminum Altetra-dist with a tetrahedral coordination. The lowered Altetra-frame concentration accounted for the decrease in Brønsted acidity for the phosphorus-modified zeolites. High-temperature calcination of P HZSM-5 resulted in formation of occluded condensed polyphosphates and Al–O–P complexes (Altetra-dist–O–P and Alocta–O–P species where Al remained at the original framework positions), which were not present in the standard HZSM-5. Washing with hot water removed most of the introduced P and partly restored framework Altetra-frame sites from these Al–O–P complexes, consequently increasing the corresponding Brønsted acidity. This could be explained by the partial reversibility of the modification process of HZSM-5 with phosphorus anions. Moreover, the treatments changed also the ratio of the framework Al sites, i.e., B1 and B2 sites. The larger extent of restoration of B2 compared to B1 acid sites increased the relative abundance of B2. The modified acidities were correlated to their catalytic performances in the methanol-to-olefins reaction.
Co-reporter:Dipl.-Chem. Jennifer Hein;Dr. Ana Hrabar;Dr. Andreas Jentys;Dr. Oliver Y. Gutiérrez;Dr. Johannes A. Lercher
ChemCatChem 2014 Volume 6( Issue 2) pp:485-499
Publication Date(Web):
DOI:10.1002/cctc.201300856

Abstract

Supported MoS2/γ-Al2O3 and Ni-MoS2/γ-Al2O3 as well as unsupported Ni-MoS2 were investigated in the hydrodenitrogenation (HDN) of quinoline in the presence of dibenzothiophene (DBT). The supported oxide catalyst precursors had a well-dispersed amorphous polymolybdate structure that led to the formation of a highly dispersed sulfide phase. In contrast, the unsupported catalyst precursor consisted of a mixture of nickel molybdate and ammonium nickel molybdate phases that formed stacked sulfide slabs after sulfidation. On all catalysts, the reaction pathway for the removal of N in quinoline HDN mainly followed the sequence quinoline1,2,3,4-tetrahydroquinolinedecahydroquinolinepropylcyclohexylaminepropylcyclohexenepropylcyclohexane. The hydrodesulfurization of DBT proceeded mainly by direct desulfurization towards biphenyl. For both processes, the activity increased in the order MoS2/γ-Al2O3<Ni-MoS2/unsupported<Ni-MoS2/γ-Al2O3. The promotion of the MoS2 phase with Ni enhances the activity of the unsupported catalyst to a greater extent than the supported one. However, the multiply stacked unsupported Ni-MoS2 exhibited lower rates than Ni-MoS2/γ-Al2O3 because of its lower dispersion.

Co-reporter:Christian A. Gärtner;Andre C. van Veen
Topics in Catalysis 2014 Volume 57( Issue 14-16) pp:1236-1247
Publication Date(Web):2014 September
DOI:10.1007/s11244-014-0277-5
Binary, ternary and quaternary molten eutectic alkali chloride catalysts, supported on mildly redox active oxides, were investigated for the oxidative dehydrogenation of ethane. The influence of different support oxides, on the catalytic performance, as well as that of different anions (bromide vs. chloride) and cations in a chloride eutectic system were studied. Metal oxides which react with chlorides are not suitable and lead to substantial deactivation. Especially supports forming volatile chlorides induce irreversible chloride depletion. Bromides catalyze oxidative dehydrogenation of ethane with higher rates, but lower olefin selectivities, highlighting the similarities and differences of Cl− and Br− in the redox cycle. Two catalysts were identified having olefin selectivities up to 98 % at 70 % ethane conversion, which ranges among the highest selectivities reported for ethane ODH.
Co-reporter:Chen Zhao, Thomas Brück and Johannes A. Lercher  
Green Chemistry 2013 vol. 15(Issue 7) pp:1720-1739
Publication Date(Web):14 May 2013
DOI:10.1039/C3GC40558C
Microalgae are high potential raw biomass material for triglyceride feedstock, due to their high oil content and rapid growth rate, and because algae cultivation does not compete with edible food on arable land. This review addresses first the microalgae cultivation with an overview of the productivity and growth of microalgae, the recovery of lipids from the microalgae, and chemical compositions of microalgae biomass and microalgal oil. Second, three basic approaches are discussed to downstream processing for the production of green gasoline and diesel hydrocarbons from microalgae oil, including cracking with zeolite, hydrotreating with supported sulfided catalysts and hydrodeoxygenation with non-sulfide metal catalysts. For the triglyceride derived bio-fuels, only “drop-in” gasoline and diesel range components are discussed in this review.
Co-reporter:John H. Ahn, Robin Kolvenbach, Sulaiman S. Al-Khattaf, Andreas Jentys, and Johannes A. Lercher
ACS Catalysis 2013 Volume 3(Issue 5) pp:817
Publication Date(Web):March 12, 2013
DOI:10.1021/cs4000766
The reaction of toluene methylation was investigated with four acidic zeolites of different pore geometries: the medium pore zeolites H-ZSM5 and H-ZSM11 as well as the large pore zeolites H-MOR and H-BEA. The methylation, methanol consumption, light hydrocarbon formation, and disproportionation rates for the reaction of toluene, p-xylene, and 1,2,4-trimethylbenezene with methanol were determined. The products of toluene methylation (e.g., xylenes and trimethylbenzenes) were readily methylated further in both medium and large pore zeolites. A considerably higher fraction of methanol was used to form light hydrocarbons with the medium pore zeolites than with large pore zeolites. This was related to the fact that the dealkylation of light hydrocarbons from highly methylated aromatics became more favorable relative to methylation at an earlier stage, that is, after fewer methyl groups were added to the aromatic ring. Increasing the effective residence time of bulky aromatic molecules with medium pore zeolites, modified either by coating the surface with tetraethyl orthosilicate or by increasing the intracrystal pore length, converted a larger fraction of methanol to light hydrocarbons via methylation and subsequent dealkylation of light hydrocarbons.Keywords: acid chemistry in zeolites; hydrocarbon pool cycle; methanol to hydrocarbons; petrochemical upgrade; toluene methylation
Co-reporter:Hui Shi, Oliver Y. Gutiérrez, Hao Yang, Nigel D. Browning, Gary L. Haller, and Johannes A. Lercher
ACS Catalysis 2013 Volume 3(Issue 3) pp:328
Publication Date(Web):January 4, 2013
DOI:10.1021/cs300636j
Ring-opening of cyclopentane on alumina-supported Pt particles was studied as a function of Pt particle size in the presence of different Cl contents. With catalysts prepared from a Cl-free precursor, measured turnover rates increased monotonically with increasing Pt particle size (1–15 nm). On catalysts derived from a Cl-containing precursor, the turnover rates fell into two separate trends with the change of Pt particle size, depending on the extent of Cl removal by increasing thermal treatment temperature. In both cases, catalytic activity increased with increasing particle size in the examined ranges of dispersions (D = 0.7–1.0 and 0–0.6) and for both series of catalysts, the apparent activation energies were higher on large Pt particles than on small ones, with only small differences in the reaction orders for H2 and cyclopentane on particles of widely varying average sizes. Therefore, the effect of particle size on the turnover rates stems mainly from intrinsic rate constants, rather than from coverage effects. The relative adsorption coefficients of toluene and benzene indicated lower electron densities at the surface Pt atoms in the catalysts prepared from the Cl-containing precursor than in those from the Cl-free precursor. This subtle electron deficiency, which seems not to stem from the local Cl enrichment near Pt, affects both the concentration of chemisorbed hydrogen under reaction conditions and the barrier for C–C bond cleavage. The Cl postintroduced to the catalyst, in contrast, does not induce a similarly positive effect.Keywords: chloride; cyclopentane; platinum; ring-opening; structure sensitivity
Co-reporter:John H. Ahn, Robin Kolvenbach, Sulaiman S. Al-Khattaf, Andreas Jentys and Johannes A. Lercher  
Chemical Communications 2013 vol. 49(Issue 90) pp:10584-10586
Publication Date(Web):23 Sep 2013
DOI:10.1039/C3CC46197A
An increase in p-xylene selectivity was observed without losing the catalytic activity over novel mesoporous nano-sized ZSM5 crystals covered with an external SiO2 overlayer created by deposition of tetraethyl orthosilicate.
Co-reporter:Oliver Y. Gutiérrez, Yanzhe Yu, Robin Kolvenbach, Gary L. Haller and Johannes A. Lercher  
Catalysis Science & Technology 2013 vol. 3(Issue 9) pp:2365-2372
Publication Date(Web):02 Jul 2013
DOI:10.1039/C3CY00189J
Pt supported on sulfated zirconia (SZ) and amorphous silica alumina (ASA) was explored for the hydrogenation of tetralin. Pt/ASA had a higher concentration and strength of acid sites than the SZ-supported counterpart, for which the acid site concentration of the carrier decreases after catalyst synthesis due to sulfur elimination. The strong acidity of Pt/ASA caused higher deactivation due to coke deposition, while Pt/SZ was comparatively stable. Pt/ASA also exhibited higher porosity and Pt dispersion than Pt/SZ. The adsorption of CO, the XANES analysis and gravimetric sorption of benzene showed that the Pt particles on SZ were strongly electron deficient, a feature that has been speculated to be associated with the electron withdrawing effect of sulfate groups. For tetralin hydrogenation, Pt/SZ was more active than Pt/ASA above 423 K while the opposite was observed at 413 K. The apparent activation energies were 98 and 45 kJ mol−1 on Pt/SZ and Pt/ASA, respectively. Pt/ASA was more active than Pt/SZ in the presence of quinoline, while Pt/SZ retains the highest activity in the presence of dibenzothiophene (with or without quinoline). The lower apparent activation energy on Pt/ASA in the absence of quinoline or dibenzothiophene and its higher activity in the presence of quinoline were attributed to its strong Brønsted acidity, promoting the hydrogenation at the perimeter of Pt particles. The higher sulfur resistance of Pt/SZ was attributed to the electron deficiency of the supported Pt particles. In this respect, surface sulfate anions induce stronger electron deficiency in supported Pt than the acidity of the support.
Co-reporter:Stefanie Simson ; Andreas Jentys
The Journal of Physical Chemistry C 2013 Volume 117(Issue 16) pp:8161-8169
Publication Date(Web):March 28, 2013
DOI:10.1021/jp3119764
It is known that vinyl acetate monomer synthesis over bimetallic Pd1Au1 catalysts is highly structurally sensitive and that these structures are dynamically formed and disintegrated in reactant gas atmosphere. Here we show via a combination of bulk and surface sensitive methods that independent of their nominal composition, Pd/Au bimetallic particles undergo marked reconstruction and phase partition during vinyl acetate synthesis. While temperature-induced reorganization of SiO2-supported PdxAuy particles leads to all three thermodynamically metastable phases, i.e., Pd3Au, Pd1Au1, PdAu3, with mainly Au-enriched surface compositions, the reactive atmosphere induces selectively the formation of the Pd1Au1 phase and separated Pd particles. The Pd in the Pd1Au1 has higher specific activity and selectivity than the sole Pd nanoparticles, exemplifying the importance of changing its electronic nature. As the Pd1Au1 phase dominates the catalytic activity, novel tailored catalysts could be produced, limiting the presence of dispersed Pd.
Co-reporter:Dr. Oliver Y. Gutiérrez;Dr. Liangshu Zhong;Dr. Yongzhong Zhu ;Dr. Johannes A. Lercher
ChemCatChem 2013 Volume 5( Issue 11) pp:
Publication Date(Web):
DOI:10.1002/cctc.201390054
Co-reporter:Dr. Oliver Y. Gutiérrez;Dr. Liangshu Zhong;Dr. Yongzhong Zhu ;Dr. Johannes A. Lercher
ChemCatChem 2013 Volume 5( Issue 11) pp:3249-3259
Publication Date(Web):
DOI:10.1002/cctc.201300210

Abstract

The synthesis of methanethiol from CS2 catalyzed by Ni-, Co-, or K-modified MoS2/SiO2 through the sequence CS2CH3SHCH3SCH3 (DMS) has been studied. CS2 is readily hydrogenated to CH3SH, whereas CH3SH and DMS decompose to CH4 at high CS2 conversions and temperatures. Thus, very high yields of CH3SH can be obtained on all the catalysts. The addition of Co or Ni accelerates all reaction steps to allow the formation of CH3SH under relatively mild conditions. On the contrary, K suppresses the formation of byproducts but also lowers the activity. The surface reaction proceeds through a series of hydrogen-addition steps, the rate-controlling step of which is the addition of the first H atom. H2 adsorbs dissociatively, whereas CS2 adsorbs strongly in a single active site prior to hydrogenation. The addition of Ni has a stronger enhancing effect on hydrogenation and accelerates the CS bond cleavage to a smaller extent than Co. Therefore, Ni leads to optimum CH3SH yields and holds the best promise as promoter for the MoS2 phase in the synthesis of methanethiol from CS2.

Co-reporter:Wenji Song;Dr. Chen Zhao;Dr. Johannes A. Lercher
Chemistry - A European Journal 2013 Volume 19( Issue 30) pp:
Publication Date(Web):
DOI:10.1002/chem.201390112
Co-reporter:Wenji Song;Dr. Chen Zhao;Dr. Johannes A. Lercher
Chemistry - A European Journal 2013 Volume 19( Issue 30) pp:9833-9842
Publication Date(Web):
DOI:10.1002/chem.201301005

Abstract

Improved synthetic approaches for preparing small-sized Ni nanoparticles (d=3 nm) supported on HBEA zeolite have been explored and compared with the traditional impregnation method. The formation of surface nickel silicate/aluminate involved in the two precipitation processes are inferred to lead to the stronger interaction between the metal and the support. The lower Brønsted acid concentrations of these two Ni/HBEA catalysts compared with the parent zeolite caused by the partial exchange of Brønsted acid sites by Ni2+ cations do not influence the hydrodeoxygenation rates, but alter the product selectivity. Higher initial rates and higher stability have been achieved with these optimized catalysts for the hydrodeoxygenation of stearic acid and microalgae oil. Small metal particles facilitate high initial catalytic activity in the fresh sample and size uniformity ensures high catalyst stability.

Co-reporter:Dr. Baoxiang Peng;Dr. Chen Zhao;Stanislav Kasakov;Sebastian Foraita ;Dr. Johannes A. Lercher
Chemistry - A European Journal 2013 Volume 19( Issue 15) pp:
Publication Date(Web):
DOI:10.1002/chem.201390049
Co-reporter:Dipl.-Ing. MSc. Christian A. Gärtner;Dr. André C. vanVeen;Dr. Johannes A. Lercher
ChemCatChem 2013 Volume 5( Issue 11) pp:3196-3217
Publication Date(Web):
DOI:10.1002/cctc.201200966

Abstract

The increasing demand for light olefins and the changing nature of basic feedstock has stimulated substantial research activity into the development of new process routes. Steam cracking remains the most industrially relevant pathway, but other routes for light-olefin production have emerged. Fluid catalytic cracking only produces ethene in minor concentrations. Challenged by marked catalyst deactivation, in contrast, catalytic dehydrogenation ethane up opens a more selective route to ethene. The oxidative dehydrogenation (ODH) of ethane, which couples the endothermic dehydration of ethane with the strongly exothermic oxidation of hydrogen, would potentially be the most attractive alternative route because it avoids the need for excessive internal heat input, but also consumes valuable hydrogen. In this Review, the current state of the ODH of ethane is compared with other routes for light-olefin production, with a focus on the catalyst and reactor system variants. New catalyst systems and reactor designs have been developed to improve the industrial competitiveness of the ODH reaction of ethane. The current state of our fundamental understanding of the ODH of light alkanes, in particular in terms of catalyst and reactor development, is critically reviewed. The proposed mechanisms and the nature of the active site for the ODH reaction are described and discussed in detail for selected promising catalysts. The reported catalytic performance and the possible limitations of these ODH catalysts will be examined and the performance of the emerging approaches is compared to the currently practiced methods.

Co-reporter:Dr. Baoxiang Peng;Dr. Chen Zhao;Stanislav Kasakov;Sebastian Foraita ;Dr. Johannes A. Lercher
Chemistry - A European Journal 2013 Volume 19( Issue 15) pp:4732-4741
Publication Date(Web):
DOI:10.1002/chem.201203110

Abstract

The mechanism of the catalytic reduction of palmitic acid to n-pentadecane at 260 °C in the presence of hydrogen over catalysts combining multiple functions has been explored. The reaction involves rate-determining reduction of the carboxylic group of palmitic acid to give hexadecanal, which is catalyzed either solely by Ni or synergistically by Ni and the ZrO2 support. The latter route involves adsorption of the carboxylic acid group at an oxygen vacancy of ZrO2 and abstraction of the α-H with elimination of O to produce the ketene, which is in turn hydrogenated to the aldehyde over Ni sites. The aldehyde is subsequently decarbonylated to n-pentadecane on Ni. The rate of deoxygenation of palmitic acid is higher on Ni/ZrO2 than that on Ni/SiO2 or Ni/Al2O3, but is slower than that on H-zeolite-supported Ni. As the partial pressure of H2 is decreased, the overall deoxygenation rate decreases. In the absence of H2, ketonization catalyzed by ZrO2 is the dominant reaction. Pd/C favors direct decarboxylation (−CO2), while Pt/C and Raney Ni catalyze the direct decarbonylation pathway (−CO). The rate of deoxygenation of palmitic acid (in units of mmol moltotal metal−1 h−1) decreases in the sequence r(Pt black)r(Pd black)>r(Raney Ni) in the absence of H2. In situ IR spectroscopy unequivocally shows the presence of adsorbed ketene (CCO) on the surface of ZrO2 during the reaction with palmitic acid at 260 °C in the presence or absence of H2.

Co-reporter:Jeongnam Kim, Andreas Jentys, Sarah M. Maier, and Johannes A. Lercher
The Journal of Physical Chemistry C 2013 Volume 117(Issue 2) pp:986-993
Publication Date(Web):December 7, 2012
DOI:10.1021/jp309277n
Structural changes to Fe3+ cationic species in Fe-exchanged zeolite BEA during the selective catalytic reduction (SCR) of NOx with NH3 were probed by UV–vis spectroscopy. The distribution between Fe2+ and Fe3+ species was characterized by IR spectroscopy of adsorbed CO. Upon heating to 723 K, some of the Fe cations formed Fe–O–Fe bonds that underwent reversible structural transformation under NH3–SCR conditions. The in situ formed Fe oxide clusters could be dissociated to isolated Fe cations at 423 K, while at higher temperatures O-bridged Fe clusters were again formed. The structure of the Fe cluster is related to the Al distribution in the zeolite probed by Co2+ ion exchange. We propose here that two Fe cations bound within one six-membered ring containing an Al pair form hydroxylated dimeric Fe–O–Fe in the zeolite. This was supported by a structure simulation of a binuclear [HO-Fe(III)–O–Fe(III)–OH] model.
Co-reporter:Baoxiang Peng ; Xiaoguo Yuan ; Chen Zhao
Journal of the American Chemical Society 2012 Volume 134(Issue 22) pp:9400-9405
Publication Date(Web):April 30, 2012
DOI:10.1021/ja302436q
A new route to convert crude microalgae oils using ZrO2-promoted Ni catalysts into diesel-range alkanes in a cascade reaction is presented. Ni nanoparticles catalyze the selective cleavage of the C–O of fatty acid esters, leading to the hydrogenolysis of triglycerides. Hydrogenation of the resulting fatty acids to aldehydes (rate-determining step) is uniquely catalyzed via two parallel pathways, one via aldehyde formation on metallic Ni and the second via a synergistic action by Ni and ZrO2 through adsorbing the carboxylic groups at the oxygen vacancies of ZrO2 to form carboxylates and subsequently abstracting the α-hydrogen atom to produce ketene, which is in turn hydrogenated to aldehydes and decarbonylated on Ni nanoparticles.
Co-reporter:Susanne Schäfer ; Sonja A. Wyrzgol ; Roberta Caterino ; Andreas Jentys ; Sebastian J. Schoell ; Michael Hävecker ; Axel Knop-Gericke ; Johannes A. Lercher ; Ian D. Sharp ;Martin Stutzmann
Journal of the American Chemical Society 2012 Volume 134(Issue 30) pp:12528-12535
Publication Date(Web):June 27, 2012
DOI:10.1021/ja3020132
Platinum nanoparticles supported on n- and p-type gallium nitride (GaN) are investigated as novel hybrid systems for the electronic control of catalytic activity via electronic interactions with the semiconductor support. In situ oxidation and reduction were studied with high pressure photoemission spectroscopy. The experiments revealed that the underlying wide-band-gap semiconductor has a large influence on the chemical composition and oxygen affinity of supported nanoparticles under X-ray irradiation. For as-deposited Pt cuboctahedra supported on n-type GaN, a higher fraction of oxidized surface atoms was observed compared to cuboctahedral particles supported on p-type GaN. Under an oxygen atmosphere, immediate oxidation was recorded for nanoparticles on n-type GaN, whereas little oxidation was observed for nanoparticles on p-type GaN. Together, these results indicate that changes in the Pt chemical state under X-ray irradiation depend on the type of GaN doping. The strong interaction between the nanoparticles and the support is consistent with charge transfer of X-ray photogenerated free carriers at the semiconductor–nanoparticle interface and suggests that GaN is a promising wide-band-gap support material for photocatalysis and electronic control of catalysis.
Co-reporter:Jiayue He ; Chen Zhao
Journal of the American Chemical Society 2012 Volume 134(Issue 51) pp:20768-20775
Publication Date(Web):November 28, 2012
DOI:10.1021/ja309915e
A novel Ni/SiO2-catalyzed route for selective cleavage of ether bonds of (lignin-derived) aromatic ethers and hydrogenation of the oxygen-containing intermediates at 120 °C in presence of 6 bar H2 in the aqueous phase is reported. The C–O bonds of α-O-4 and β-O-4 linkages are cleaved by hydrogenolysis on Ni, while the C–O bond of the 4-O-5 linkage is cleaved via parallel hydrogenolysis and hydrolysis. The difference is attributed to the fact that the Caliphatic–OH fragments generated from hydrolysis of α-O-4 and β-O-4 linkages can undergo further hydrogenolysis, while phenol (produced by hydrolysis of the 4-O-5 linkage) is hydrogenated to produce cyclohexanol under conditions investigated. The apparent activation energies, Ea(α-O-4) < Ea(β-O-4) < Ea(4-O-5), vary proportionally with the bond dissociation energies. In the conversion of β-O-4 and 4-O-5 ether bonds, C–O bond cleavage is the rate-determining step, with the reactants competing with hydrogen for active sites, leading to a maximum reaction rate as a function of the H2 pressure. For the very fast C–O bond cleavage of the α-O-4 linkage, increasing the H2 pressure increases the rate-determining product desorption under the conditions tested.
Co-reporter:Chen Zhao, Wenji Song, and Johannes A. Lercher
ACS Catalysis 2012 Volume 2(Issue 12) pp:2714
Publication Date(Web):October 19, 2012
DOI:10.1021/cs300418a
Aqueous phase catalytic phenol hydroalkylation and hydrodeoxygenation have been explored using Pd/C combined with zeolite H-BEA and La-BEA catalysts in the presence of H2. The individual steps of phenol hydrogenation, cyclohexanol dehydration, or alkylation with phenol were individually investigated to gain insight into the relative rates in the cascade reactions of phenol hydroalkylation. The hydroalkylation rate, determined by the concentrations of phenol and cyclohexanol in phenol hydroalkylation, required the hydrogenation rate to be relatively slow. The optimized H+/Pd ratio was 21, which allowed achieving comparable cyclohexanol formation rates via phenol hydrogenation and consumption rates from alkylation with phenol in phenol hydroalkylation. La-BEA was shown to be more selective for hydroalkylation than H-BEA in combination with Pd/C, because cyclohexanol dehydration was retarded selectively compared to alkylation of phenol. This indicates that dehydration is solely catalyzed by Brønsted acid sites, while alkylation can be achieved in the presence of La3+ cations.Keywords: deoxygenation; hydroalkylation; hydrogenation; phenol; sustainable chemistry;
Co-reporter:Lay Hwa Ong, Márta Dömök, Roberta Olindo, André C. van Veen, Johannes A. Lercher
Microporous and Mesoporous Materials 2012 Volume 164() pp:9-20
Publication Date(Web):1 December 2012
DOI:10.1016/j.micromeso.2012.07.033
Two distinct locations for tetrahedral aluminum in HZSM-5 have been identified, showing bond angles of 150° and 155° T–O–T, respectively. The former site is more abundant and has been indirectly associated with aluminum in isolated positions. Upon steaming its concentration decreases, following a first order rate law and leading to the formation of tetrahedrally coordinated extra-lattice aluminum as well as to invisible extra-lattice aluminum. The latter is speculated to be also tetrahedrally coordinated and kinetically linked to the visible portion of the extra-lattice aluminum. Both types of extra-lattice aluminum neutralize lattice charge and lead to a decrease of the concentration of bridging Si–OH–Al groups, which is initially more pronounced than the loss of lattice aluminum. With steaming duration the concentration of Brønsted acid sites stabilizes indicating that the extra-lattice aluminum atoms begin to form larger clusters at a rate equivalent to the rate of dealumination. The lattice aluminum with the more obtuse T–O–T angle is stable under the steaming conditions chosen. As its concentration is nearly equivalent to the concentration of aluminum sites sufficiently close to exchange Co2+ ions, it is inferred that aluminum resisting dealumination constitutes these sites.Graphical abstractHighlights► Two tetrahedral Al species have been identified by 27Al MAS NMR at 54.4 and 57 ppm. ► The species at 54.4 ppm corresponding to the more obtuse T–O–T angle (155°) was stable during steaming. ► The concentration of the steam-stable tetrahedral Al was identical with the Co2+ exchange capacity. ► The tetrahedral and invisible extra-lattice Al species generated during steaming neutralize potential zeolite acid sites.
Co-reporter:Sabine Scholz;Hui Shi
Topics in Catalysis 2012 Volume 55( Issue 11-13) pp:800-810
Publication Date(Web):2012 August
DOI:10.1007/s11244-012-9852-9
Hierarchically structured macroscopic Pt-silica spheres were prepared in a one-step procedure using emulsion based sol–gel processing. The composition of the precursor solution was varied in order to study the influence of the Pt compound, the surfactant and the preparation method on the Pt particle size and concentration. Platinum was immobilized by coordination with an aminosilane and the nanoparticles were stabilized using different triblock copolymers. For removal of the surfactant, calcination as well as solvent extraction was used and the co-condensed samples were compared with silica spheres functionalized by ion-exchange. The activities of different catalysts were investigated in the methylcyclopentane ring opening reaction, which demonstrated a beneficial effect of the macroscopic shape and hierarchical structure of the support material. The simultaneous structural control over several scales of dimension from the macroscopic shape of the support to the nanometer size of the Pt nanoparticles allows the easy and highly reproducible synthesis of tailor-made catalysts.
Co-reporter:Baoxiang Peng;Yuan Yao;Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie 2012 Volume 124( Issue 9) pp:2114-2117
Publication Date(Web):
DOI:10.1002/ange.201106243
Co-reporter:Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie International Edition 2012 Volume 51( Issue 24) pp:
Publication Date(Web):
DOI:10.1002/anie.201201227
Co-reporter:Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie International Edition 2012 Volume 51( Issue 24) pp:5935-5940
Publication Date(Web):
DOI:10.1002/anie.201108306
Co-reporter:Baoxiang Peng;Yuan Yao;Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie International Edition 2012 Volume 51( Issue 9) pp:
Publication Date(Web):
DOI:10.1002/anie.201108371
Co-reporter:Baoxiang Peng;Yuan Yao;Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie International Edition 2012 Volume 51( Issue 9) pp:2072-2075
Publication Date(Web):
DOI:10.1002/anie.201106243
Co-reporter:Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie 2012 Volume 124( Issue 24) pp:
Publication Date(Web):
DOI:10.1002/ange.201201227
Co-reporter:Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie 2012 Volume 124( Issue 24) pp:6037-6042
Publication Date(Web):
DOI:10.1002/ange.201108306
Co-reporter:Baoxiang Peng;Yuan Yao;Dr. Chen Zhao ;Dr. Johannes A. Lercher
Angewandte Chemie 2012 Volume 124( Issue 9) pp:
Publication Date(Web):
DOI:10.1002/ange.201108371
Co-reporter:Dr. Chen Zhao ;Dr. Johannes A. Lercher
ChemCatChem 2012 Volume 4( Issue 1) pp:64-68
Publication Date(Web):
DOI:10.1002/cctc.201100273
Co-reporter:Oliver Y. Gutiérrez, Christoph Kaufmann, and Johannes A. Lercher
ACS Catalysis 2011 Volume 1(Issue 11) pp:1595
Publication Date(Web):October 6, 2011
DOI:10.1021/cs200455k
The catalytic properties of a series of (Co)K-promoted Mo sulfide catalysts supported on SiO2 were explored in the synthesis of methanethiol from carbonyl sulfide (COS) and CS2. MoS2 was very active for the conversion of COS, but allowed only low yields of CH3SH because of the parallel decomposition of COS to CO and H2S and the reduction of CH3SH to CH4. CS2, on the other hand, was completely converted to CH3SH with high yield below 570 K on MoS2. The formation of CH4, however, dramatically decreased the yield of CH3SH above 570 K. The addition of K+ cations decreased the conversion of both reactants, but also reduced the rate of decomposition/reduction reactions. The doubly promoted CoK-Mo catalyst led to the highest conversions with moderate to high yields of methanethiol. We conclude that the addition of K+ cations generates very weak adsorption sites, suppressing so the C–S bond cleavage. These sites catalyze, however, COS disproportionation. Accessible Co and Mo sites are part of the active sites for all reactions observed. All catalytic active sites are concluded to be on the edge of MoS2 slabs.Keywords: carbon disulfide; carbonyl sulfide; K−Co−Mo sulfide catalyst; methanethiol;
Co-reporter:Michael Salzinger and Johannes A. Lercher  
Green Chemistry 2011 vol. 13(Issue 1) pp:149-155
Publication Date(Web):30 Nov 2010
DOI:10.1039/C0GC00428F
The reaction network and mechanism of the synthesis of methylenedianiline (MDA) from the condensation product of aniline and formaldehyde (aminal) on microporous acidic materials has been elucidated. The first step of the reaction, the decomposition of the aminal to N-benzylanilines, is limited by film diffusion, and the second and significantly slower step, the acid catalyzed rearrangement of these intermediates to MDA, is controlled by microkinetics on mesoporous dealuminated Y-type zeolites. This second step of the reaction network is limited by pore diffusion on zeolite BEA as an example for non-mesoporous materials. Based on time-concentration profiles, we were able to determine the reaction orders of the initial decomposition of the aminal to one and two for the following rearrangement of para-aminobenylaniline to 4,4′-MDA. From the kinetic data we deduced an SN2-type reaction mechanism and a complex reaction network, which is able to simulate the observed concentration profiles. The aniline to formaldehyde ratio in the starting mixture had a negligible influence on the final product distribution.
Co-reporter:Sabine Scholz and Johannes A. Lercher
Chemistry of Materials 2011 Volume 23(Issue 8) pp:2091
Publication Date(Web):March 29, 2011
DOI:10.1021/cm1033029
Tailored hierarchically structured (organo) silica beads with particle diameters between 0.3 and 2.5 mm have been prepared using an emulsion based sol−gel approach. Poly(propylene oxide)-poly(ethylene oxide) triblock copolymers as surfactants, aniline or 1-butanol as solvent and different silanes such as tetraethyl silicate, [3-(2-aminoethylamino)propyl]trimethoxysilane and phenyltrimethoxysilane are used to direct morphology and porosity. The silica spheres provide specific surface areas up to 900 m2/g and adjustable porosity in the meso- and macroporous region. The texture includes particulate structures with controllable size between 80 nm and 2.5 μm of the silica particles on the inside of the spheres and a macroporous foam-like shell with macropores in the range between 0.2 and 4.0 μm and a thickness of the shell between 0.6 and 44 μm. The high variability of this preparation method and the simultaneous control of the nano- to millimeter scale structure including the porosity as well as the morphology allows one to design tailor-made adsorbents and catalysts.Keywords: emulsion-based sol−gel synthesis; hierarchical structure; macroporous shell; meso/microporous core; tunable millimeter sized (organo) silica spheres;
Co-reporter:Manuela C. I. Bezen, Cornelia Breitkopf, and Johannes A. Lercher
ACS Catalysis 2011 Volume 1(Issue 10) pp:1384
Publication Date(Web):September 1, 2011
DOI:10.1021/cs200175x
Mg/Al mixed oxides derived from layered double hydroxide precursors were modified with fluoride anions to probe their influence on acid–base and catalytic properties. Fluoride anions were incorporated by treatment with aqueous NH4F. The structure of the precursors and the modified materials were analyzed by XRD, 27Al NMR, and 19F MAS NMR. The modification of Mg/Al mixed oxides led to the incorporation of fluoride anions in various configurations. Overall, the concentration of acid sites was enhanced by this modification while it decreased the concentration of the basic sites. The rate of dehydrogenation of propan-2-ol drastically increased in the presence of F– anions, but the rate of dehydration was hardly influenced. This indicates that the fluoride anions improve the ability of the mixed oxides to abstract protons from polar molecules without increasing the overall base strength of the mixed oxide.Keywords: 27Al/19F MAS NMR; dehydrogenation of propan-2-ol; F− modification; Mg/Al mixed oxides; XRD;
Co-reporter:Tobias Förster, Sabine Scholz, Yongzhong Zhu, Johannes A. Lercher
Microporous and Mesoporous Materials 2011 Volume 142(2–3) pp:464-472
Publication Date(Web):July 2011
DOI:10.1016/j.micromeso.2010.12.032
Hydrophobic mesoporous silica spheres containing transition metals coordinatively bound to aminosilane groups are synthesized in a one-step procedure from functionalized silicon alkoxides. In situ IR spectroscopy shows the presence of amino groups in the solid. These amino groups are critical to bind transition metals such as Co, Cu, Fe, Mn or V by complexation. The accessibility of the transition metals is shown by IR spectra of adsorbed CO. The method described provides a scalable approach to synthesize tailored transition metal containing structured micro and mesoporous organofunctionalized silica spheres, which show great potential for catalytic applications.Graphical abstractHydrophobic polysiloxane spheres in the millimeter range have been successfully synthesized from silicon alkoxide precursors. Via amino functions transition metal cations are coordinated to the support for potential catalytic applications.Research highlights► One step synthesis of macroscopic transition metal polysiloxane beads. ► Management of surface hydrophobicity by variation of precursors. ► Accessible transition metals show great potential for catalytic applications.
Co-reporter:Sabine Scholz, Simon R. Bare, Shelly D. Kelly, Johannes A. Lercher
Microporous and Mesoporous Materials 2011 Volume 146(1–3) pp:18-27
Publication Date(Web):December 2011
DOI:10.1016/j.micromeso.2011.04.036
Novel hierarchically structured macroscopic silica spheres have been synthesized using a biphasic oil-in-water system. The spheres are prepared in a one-step process by injection of the precursor solution into water. This leads to controlled hydrolysis and condensation of the silanes. The precursor solution contained three different silanes, a block copolymer acting as surfactant and a solvent. The spheres have a hierarchically layered morphology, high mechanical stability, specific surface areas up to 700 m2/g with mesopores in the range of 9–27 nm. The synthesis of such materials with a simultaneous control of the macroscopic and microscopic structure allows facile implementation in separation and catalysis.Graphical abstractHighlights► One-step emulsion based synthesis of millimeter-sized silica spheres. ► Hierarchical core-shell morphology. ► Controlled macro- and mesopores via specific surfactant molecules.
Co-reporter:Sarah M. Maier, Andreas Jentys, Ezzeldin Metwalli, Peter Müller-Buschbaum, and Johannes A. Lercher
The Journal of Physical Chemistry Letters 2011 Volume 2(Issue 9) pp:950-955
Publication Date(Web):April 8, 2011
DOI:10.1021/jz200313c
The nature and oxidation state of iron species in Fe-exchanged BEA zeolites treated in synthetic air or nitrogen were determined by a combination of Mössbauer and X-ray absorption spectroscopy. The linear correlation between the edge energy of the X-ray absorption near edge structure (XANES) and the oxidation state determined by Mössbauer spectroscopy allowed determining the fraction of Fe2+ in situ. The distribution of Fe2+ and Fe3+ in the catalysts depends on the Fe concentration and the conditions of the thermal treatment. It is possible to stabilize isolated Fe2+ cations under ambient atmosphere in the zeolite pores, while FeBEA catalysts show a temperature-dependent oxidation and reduction of the active Fe species during the selective catalytic reduction of nitrogen oxides by NH3 (NH3-SCR), reflecting the equilibrium for NO oxidation.Keywords: determination of oxidation states; FeBEA; Mössbauer spectroscopy; NH3-SCR; XANES;
Co-reporter:Florian Schüßler ; Evgeny A. Pidko ; Robin Kolvenbach ; Carsten Sievers ; Emiel J. M. Hensen ; Rutger A. van Santen
The Journal of Physical Chemistry C 2011 Volume 115(Issue 44) pp:21763-21776
Publication Date(Web):September 23, 2011
DOI:10.1021/jp205771e
The nature, concentration, and location of cationic lanthanum species in faujasite-type zeolites (zeolite X, Y and ultrastabilized Y) have been studied in order to understand better their role in hydrocarbon activation. By combining detailed physicochemical characterization and DFT calculations, we demonstrated that lanthanum cations are predominantly stabilized within sodalite cages in the form of multinuclear OH-bridged lanthanum clusters or as monomeric La3+ at the SI sites. In high-silica faujasites (Si/Al = 4), monomeric [La(OH)]2+ and [La(OH)2]+ species were only found in low concentrations at SII sites in the supercages, whereas the dominant part of La3+ is present as multinuclear OH-bridged cationic aggregates within the sodalite cages. Similarly, in the low-silica (Si/Al = 1.2) La–X zeolite, the SI′ sites were populated by hydroxylated La species in the form of OH-bridged bi- and trinuclear clusters. In this case, the substantial repulsion between the La3+ cations confined within the small sodalite cages induces the migration of La3+ cations into the supercage SII sites. The uniquely strong polarization of hydrocarbon molecules sorbed in La–X zeolites is caused solely by the interaction with the accessible isolated La3+ cations.
Co-reporter:Sarah M. Maier ; Andreas Jentys
The Journal of Physical Chemistry C 2011 Volume 115(Issue 16) pp:8005-8013
Publication Date(Web):April 1, 2011
DOI:10.1021/jp108338g
The kinetics and impact of steaming on the acid site concentration of zeolite beta were studied by 1H, 27Al, and 29Si MAS NMR spectroscopy as well as IR spectroscopy of adsorbed pyridine and temperature-programmed desorption of NH3. The main effects of steaming were the dealumination of the T3−T9 sites, the formation and migration of extra-framework Al species, and the healing of defect sites by condensation of silanol groups. These effects took place mainly in the first five hours of steaming, while after 14 h of steaming the system appeared to be stabilized. The concentration of framework Al atoms detected by 27Al MAS NMR spectroscopy is significantly higher than the concentration of Brønsted acid sites determined by 1H MAS NMR spectroscopy as well as by the sorption of basic probe molecules such as NH3 and pyridine. This shows conclusively that extra-framework Al oxide/hydroxide species act as cations balancing the framework charge. The concentration of extra-framework Al atoms matches the discrepancy between the concentration of framework Al atoms and the concentration of Brønsted acid sites indicating that each charge balancing entity contains only one aluminum.
Co-reporter:O. C. Gobin ; S. J. Reitmeier ; A. Jentys ;J. A. Lercher
The Journal of Physical Chemistry C 2011 Volume 115(Issue 4) pp:1171-1179
Publication Date(Web):December 21, 2010
DOI:10.1021/jp106474x
The role of the surface modification on the transport properties of nanosized HZSM-5 particles was investigated by the analysis of the sorption kinetics of hexane isomers. The rate of diffusion is enhanced by this modification, if the initial adsorption is the limiting step, i.e., for n-hexane, 2-methylpentane, and 3-methylpentane. If the intracrystalline transport is the rate-determining step, i.e., for 2,2-dimethylbutane, the surface modification leads to a significantly decreased sorption rate. This reduction in the transport rate is caused by a lower pre-exponential factor; i.e., it is entropically induced. Due to the surface modification, a certain fraction of the pores is completely blocked for the sterically demanding molecules reducing the probability of entering the pores. The results show that surface modification of microporous materials allows selectively enhancing or decreasing the transport rate of defined systems.
Co-reporter:Manuela C.I.Bezen;Dr. Cornelia Breitkopf;Nadia ElKolli;Jean-Marc Krafft;Dr. Catherine Louis;Dr. Johannes A. Lercher
Chemistry - A European Journal 2011 Volume 17( Issue 25) pp:7095-7104
Publication Date(Web):
DOI:10.1002/chem.201002011

Abstract

Au supported on CeO2 prepared by deposition–precipitation with urea leads to a basic catalyst. Au acts in two ways as surface modifier. First, Au selectively interacts with Ce4+ cations by either blocking access to or reducing Ce4+ to Ce3+. Second, the resulting Au atoms (presumably as Au+ ions) act as soft, weak Lewis acid sites stabilizing carbanion intermediates and enhancing hydride abstraction in the dehydrogenation of alcohols. In consequence, the thus-synthesized basic catalyst catalyzes the dehydrogenation of propan-2-ol to acetone with high efficiency and without notable deactivation. Additionally, the dehydration pathway of propan-2-ol is eliminated, as Au also quantitatively blocks access to strongly acidic Ce4+ ions or reduces them to Ce3+.

Co-reporter:Felix Eckstorff, Yongzhong Zhu, Robert Maurer, Thomas E. Müller, Sabine Scholz, Johannes A. Lercher
Polymer 2011 Volume 52(Issue 12) pp:2492-2498
Publication Date(Web):26 May 2011
DOI:10.1016/j.polymer.2011.01.050
Linked octahedral silsesquioxanes and spherosilicates offer a unique approach to tailor materials with low dielectric constants. The key lies in the design of the linking structures, which is determined by the nature of the hydrolyzable alkoxysilyl groups in the monomer. Thus, the molecular structure of the monomers influences structural, dielectric, and mechanical properties of the spin-coated polymer films via the number of alkoxysilyl groups introduced. Dielectric constants and hardness can be tuned in the range k = 2.4–3.0 and 0.20–0.85 GPa, respectively. With the use of a porogen, a dielectric constant as low as 1.9 is achieved.
Co-reporter:Dr. Oliver Y. Gutiérrez;Christoph Kaufmann;Dr. Johannes A. Lercher
ChemCatChem 2011 Volume 3( Issue 9) pp:1480-1490
Publication Date(Web):
DOI:10.1002/cctc.201100124

Abstract

Potassium-doped MoS2 catalysts supported on Al2O3 were synthesized, characterized by using atomic absorption spectroscopy, N2 physisorption, NO adsorption, X-ray diffraction, temperature-programmed sulfidation, and Raman spectroscopy, and tested in the synthesis of methanethiol from carbonyl sulfide (COS) and H2. The results revealed that two phases, pure MoS2 and potassium-decorated MoS2 (formed at high potassium loadings), were present in the active catalysts. The main effect of potassium during sulfidation and during the catalytic reaction was to increase the mobility of surface oxygen or sulfur atoms. Thus, potassium promoted the disproportionation of COS to CO2 and CS2 and the production of CO from CO2. Additionally, potassium cations hindered the reductive decomposition of COS to CO and H2S and the hydrogenolysis of methanethiol to methane. Mars–van Krevelen-type mechanisms were proposed to explain the disproportionation of COS on alumina and on the MoS2 phases. The catalytic site in the potassium-decorated MoS2 phase was proposed to include a potassium cation as adsorption site.

Co-reporter:Dr. Virginia. M. Roberts;Valentin Stein;Thomas Reiner; Angeliki Lemonidou ;Dr. Xuebing Li; Johannes A. Lercher
Chemistry - A European Journal 2011 Volume 17( Issue 21) pp:5939-5948
Publication Date(Web):
DOI:10.1002/chem.201002438

Abstract

The products of base-catalyzed liquid-phase hydrolysis of lignin depend markedly on the operating conditions. By varying temperature, pressure, catalyst concentration, and residence time, the yield of monomers and oligomers from depolymerized lignin can be adjusted. It is shown that monomers of phenolic derivatives are the only primary products of base-catalyzed hydrolysis and that oligomers form as secondary products. Oligomerization and polymerization of these highly reactive products, however, limit the amount of obtainable product oil containing low-molecular-weight phenolic products. Therefore, inhibition of concurrent oligomerization and polymerization reactions during hydrothermal lignin depolymerization is important to enhance product yields. Applying boric acid as a capping agent to suppress addition and condensation reactions of initially formed products is presented as a successful approach in this direction. Combination of base-catalyzed lignin hydrolysis with addition of boric acid protecting agent shifts the product distribution to lower molecular weight compounds and increases product yields beyond 85 %.

Co-reporter:F. N. Naraschewski;A. Jentys;J. A. Lercher
Topics in Catalysis 2011 Volume 54( Issue 10-12) pp:
Publication Date(Web):2011 July
DOI:10.1007/s11244-011-9686-x
The selective oxidation of propane to acrylic acid is studied over a series of nearly pure M1-phase MoVTeNbOx catalysts. Quantitative analysis of the reaction network shows that the ratio of the rate constants for propane oxidative dehydrogenation to propene and for the further oxidation of propene is constant. The rates towards acrolein and acetone, however, vary subtly with the concentration of vanadium and the location of its substitution. The reaction of acrolein to acetic acid and carbon oxides, associated with accessible metal cations, contributes two-thirds towards the non-selective pathway. The other third is associated with acetone formation. Vanadium is first substituted selectively at sites that are inactive for propane activation. Depending on the selectivity of this substitution two groups of materials have been identified, which show a distinctly different dependence on the concentration of vanadium. Statistic distribution of vanadium in the M1 phase appears to be the most promising strategy to improve the performance of MoVTeNbOx catalysts for a given vanadium concentration.
Co-reporter:Chen Zhao, Yuan Kou, Angeliki A. Lemonidou, Xuebing Li and Johannes A. Lercher  
Chemical Communications 2010 vol. 46(Issue 3) pp:412-414
Publication Date(Web):18 Nov 2009
DOI:10.1039/B916822B
A simple, green, cost- and energy-efficient route for converting phenolic components in bio-oil to hydrocarbons and methanol has been developed, with nearly 100% yields. In the heterogeneous catalysts, RANEY® Ni acts as the hydrogenation catalyst and Nafion/SiO2 acts as the Brønsted solid acid for hydrolysis and dehydration.
Co-reporter:Sonja A. Wyrzgol, Susanne Schäfer, Sungsik Lee, Byeongdu Lee, Marcel Di Vece, Xuebing Li, Sönke Seifert, Randall E. Winans, Martin Stutzmann, Johannes A. Lercher and Stefan Vajda  
Physical Chemistry Chemical Physics 2010 vol. 12(Issue 21) pp:5585-5595
Publication Date(Web):27 Apr 2010
DOI:10.1039/B926493K
The preparation, characterization and catalytic reactivity of a GaN supported Pt catalyst in the hydrogenation of ethene are presented in this feature article, highlighting the use of in situ characterization of the material properties during sample handling and catalysis by combining temperature programmed reaction with in situ grazing incidence small-angle X-ray scattering and X-ray absorption spectroscopy. The catalysts are found to be sintering resistant at elevated temperatures as well as during reduction and hydrogenation reactions. In contrast to Pt particles of approximately 7 nm diameter, smaller particles of 1.8 nm in size are found to dynamically adapt their shape and oxidation state to the changes in the reaction environment. These smaller Pt particles also showed an initial deactivation in ethene hydrogenation, which is paralleled by the change in the particle shape. The subtle temperature-dependent X-ray absorbance of the 1.8 nm sized Pt particles indicates that subtle variations in the electronic structure induced by the state of reduction by electron tunnelling over the Schottky barrier between the Pt particles and the GaN support can be monitored.
Co-reporter:M. F. Williams ; B. Fonfé ; A. Jentys ; C. Breitkopf ; J. A. R. van Veen ;J. A. Lercher
The Journal of Physical Chemistry C 2010 Volume 114(Issue 34) pp:14532-14541
Publication Date(Web):August 5, 2010
DOI:10.1021/jp104286s
The catalytic hydrogenation of aromatic molecules, such as tetralin, on platinum supported on amorphous silica−alumina is significantly reduced by the presence of quinoline and/or dibenzothiophene compared with the reactions with pure tetralin. Quinoline neutralizes acid sites and weakens the positive effect of the acidic carrier on platinum, whereas dibenzothiophene and sulfur, resulting from dibenzothiophene hydrogenolysis, poisons selectively the metal sites. In the latter case, acid sites at the perimeter of the supported metal particles serve to bind the substrates more efficiently than Lewis acid sites and are crucial for maintaining catalytic activity. The presence of both poisons drastically reduces this pathway, lowering the catalytic activity far beyond the additive behavior.
Co-reporter:Christoph Kaufmann;Oliver Y. Gutiérrez
Research on Chemical Intermediates 2010 Volume 36( Issue 2) pp:211-225
Publication Date(Web):2010 March
DOI:10.1007/s11164-010-0131-8
The synthesis of COS from CO, CO2 and liquid sulfur in the presence and absence of hydrogen was explored. The reaction of H2 with liquid sulfur produced H2S and polysulfanes, which increase the reactivity of liquid sulfur and provide alternative complementary reaction routes for the formation of COS. The reaction from CO2 proceeds by forming CO as intermediate. Elevated pressure favors formation of COS from both carbon oxides due to the increasing residence time and the saturation of gases in the liquid. Above 350 °C, the solubility of H2S in sulfur and the hydrogenation of COS limit the conversion of CO. The approach provides a highly efficient method for the preparation of COS under mild reaction conditions, without using a catalyst or water adsorbents.
Co-reporter:Nianhua Xue, Roberta Olindo, and Johannes A. Lercher
The Journal of Physical Chemistry C 2010 Volume 114(Issue 37) pp:15763-15770
Publication Date(Web):2017-2-22
DOI:10.1021/jp106621d
The individual and combined effect of phosphoric acid modification, the addition of an alumina-based binder, and steaming on the acid−base and catalytic properties of HZSM-5 have been investigated. Whereas all three processes reduce the concentration of Brønsted acid sites, the underlying mechanisms and the extent of the effects vary. Upon addition of an alumina-based binder or during steaming of the parent zeolite dealumination is the sole process occurring. Impregnation of HZSM-5 with phosphoric acid leads to hydrolysis of framework aluminum and to the condensation of Brønsted acid sites with POH groups. This latter process can be reversed in the presence of an aqueous phase. In the forming process, this leads to the reversible formation of phosphate anions and phosphoric acid and the redistribution of phosphate species between the zeolite and the alumina binder. This mobility is mitigated by introducing phosphate species to the (alumina) binder indicating that in aqueous phase equilibrium between phosphate ions and phosphoric acid the zeolite and the (alumina) binder exists. Whatever the first step of modification of the parent material is (doping with H3PO4 or addition of either alumina or phosphated alumina binder), steaming at elevated temperatures essentially reduces and equalizes the concentration of Brønsted acid sites. For all samples derived from one common parent zeolite, the rate of protolytic cracking of propane depends linearly on the concentration of strong Brønsted acid sites; that is, all modified materials have the same turnover frequency. This turnover frequency is notably higher than that of the parent material.
Co-reporter:Richard Knapp, Sonja A. Wyrzgol, Manuela Reichelt, Tobias Hammer, Harald Morgner, Thomas E. Müller and Johannes A. Lercher
The Journal of Physical Chemistry C 2010 Volume 114(Issue 32) pp:13722-13729
Publication Date(Web):July 26, 2010
DOI:10.1021/jp103250k
Ionic liquids such as 1-butyl-2,3-dimethyl-imidazolium trifluoromethane sulfonate stabilize complex structures of polyvinylpyrrolidone protected nanoparticles on their surface. The polyvinylpyrrolidone shell around the Pt nanoparticles remains intact, but the ionic liquid partly penetrates the protective layer and interacts with the metal. This leads to a pronounced increase of the viscosity of the ionic liquid, as observed also with metal complexes. The protected particles form small islands of aggregated moieties that are connected by interacting polymer strands, a process enhanced by diluting the sample in methanol during preparation. Some of these ensembles segregate at the surface protruding significantly out from the bulk ionic liquid. The outermost layer of the surface, however, consists of the ionic liquid with the cation being on top and the SO3 group of the anion underneath being oriented toward the bulk.
Co-reporter:Dr. Virginia M. Roberts;Dr. Richard T. Knapp;Dr. Xuebing Li ;Dr. Johannes A. Lercher
ChemCatChem 2010 Volume 2( Issue 11) pp:1407-1410
Publication Date(Web):
DOI:10.1002/cctc.201000181
Co-reporter:Richard Knapp, Andreas Jentys and Johannes A. Lercher  
Green Chemistry 2009 vol. 11(Issue 5) pp:656-661
Publication Date(Web):18 Mar 2009
DOI:10.1039/B901318K
Silica supported platinum catalysts coated with a thin film of 1-butyl-2,3-dimethyl-imidazolium trifluoromethane sulfonate (BDiMIm) were investigated with respect to the interaction of the ionic liquid with the oxide support and the metal clusters. IR, inelastic neutron scattering and NMR spectroscopy indicate that the vibrations of the imidazolium ring of ionic liquid are less restricted when supported on SiO2, while the viscosity of the supported ionic liquid increased. The presence of Pt particles enhances the electron density of the ionic liquid at the nitrogen atom inducing higher basicity. The coverage of the catalyst surface and the metal particles by the ionic liquid protects the metal against oxidation. The catalysts are active and stable for hydrogenation of ethene.
Co-reporter:Elvira Peringer;Michael Salzinger;Markus Hutt
Topics in Catalysis 2009 Volume 52( Issue 9) pp:
Publication Date(Web):2009 August
DOI:10.1007/s11244-009-9265-6
Mixtures of LaOCl and LaCl3 are promising catalysts for oxidative chlorination of methane to methyl chloride. The influence of metal dopants such as Co, Ni and Ce, which form stable chlorides under anticipated reaction conditions, on physicochemical and catalytic properties was explored. The presence of markedly redox-active dopants such as cobalt and cerium lead to a higher rate of methane conversion. However, the formed methyl chloride is strongly adsorbed and directly oxidized to CO leading to low methyl chloride selectivity. Doping with nickel weakens, in contrast, the interaction with methyl chloride leading to high methyl chloride selectivity.
Co-reporter:Chen Zhao;Yuan Kou Dr.;AngelikiA. Lemonidou Dr.;Xuebing Li Dr.;JohannesA. Lercher Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 22) pp:3987-3990
Publication Date(Web):
DOI:10.1002/anie.200900404
Co-reporter:S. J. Reitmeier, O. C. Gobin, A. Jentys and J. A. Lercher
The Journal of Physical Chemistry C 2009 Volume 113(Issue 34) pp:15355-15363
Publication Date(Web):July 23, 2009
DOI:10.1021/jp905307b
Fast, time-resolved infrared spectroscopy was utilized to determine the transport and sorption properties of benzene, toluene, and p-xylene on a series of zeolites with increasing degree of surface modification. Postsynthetic modification of H-ZSM5 with a mesoporous silica overlayer generates novel hierarchical materials, which significantly increase the sorption rates for benzene and decrease it gradually for alkyl-substituted aromatic molecules with increasing radius of gyration of the aromatic molecule. The ratio between the sorption rates of benzene and p-xylene increased from an initial value of 4.3 to about 27 after modification. The reason for the enforced differentiation is the combination of the intrinsic size exclusion properties of the zeolite micropores and the variation of the sticking probability of the molecules on the modified surface, which enhanced the mass transfer into the porous overlayer. The size of the overlayer pores and the radius of gyration of the sorbate were identified as the parameters determining the sorption rates. Our results highlight that hierarchical composites generated by deposition of highly porous silica overlayers on microporous materials allow tailoring the separation properties of porous materials and allow introducing a new concept to kinetically separate molecules of identical minimum (kinetic) diameters.
Co-reporter:Yongzhong Zhu;Thomas E. Müller
Advanced Functional Materials 2008 Volume 18( Issue 21) pp:3427-3433
Publication Date(Web):
DOI:10.1002/adfm.200701394

Abstract

Composite films with low dielectric constants (k) containing micro- and mesopores are synthesized from precursor solutions for the preparation of mesoporous silica and ethanolic suspensions of silicalite-1 nanoparticles. The material contains silicalite-1 nanoparticles (include nanocrystals and nanoslabs/intermediates) embedded in a randomly oriented matrix of highly porous mesoporous silica. Micropores result from the incorporated silicalite-1 nanoparticles, while decomposition of the porogen F127 leads to additional mesopores. The porosity of the composite films increases from 9 to 60% with the increase in porogen loading, while in parallel the elastic modulus and hardness decrease. The elastic moduli of the films are in the range of 13–20 GPa. Hydrophobic surfaces of the composite films are obtained by introducing methyl triethoxysilane during the preparation of both precursor solutions, leading to the incorporation of CH3 groups in the final composite films. These methyl groups are stable up to at least 500 °C. A low k value of approximately 2 is observed for films cured at 400 °C in N2 flow, which is ideal for removing templates without decomposing methyl groups. Due to the intrinsic hydrophobicity of the material, post-silylation is not required rendering the composite films attractive candidates for future low k materials.

Co-reporter:Chinthala Praveen Kumar;Stefan Gaab;Thomas E. Müller
Topics in Catalysis 2008 Volume 50( Issue 1-4) pp:156-167
Publication Date(Web):2008 November
DOI:10.1007/s11244-008-9102-3
Potential and limitations of molten alkali metal (Li, Na, and K) chlorides supported on Dy2O3/MgO were explored for the oxidative dehydrogenation of lower alkanes, such as ethane and propane. The catalysts have high activity and selectivity to olefins compared to conventional catalysts. Optimum performance is obtained with catalysts on which the alkali metal chloride phase is molten under reaction conditions. Lower chloride melting point correlates with higher selectivity. The high selectivity to ethene or propene is attributed to the high mobility of cations and anions, which facilitates desorption of alkene (limiting further oxidation) and the generation of spatially isolated hypochloride anions acting as the active sites for the primary C–H bond activation.
Co-reporter:Xuebing Li;Katsutoshi Nagaoka;Laurent J. Simon;Roberta Olindo
Catalysis Letters 2007 Volume 113( Issue 1-2) pp:34-40
Publication Date(Web):2007 January
DOI:10.1007/s10562-006-9005-5
Calcination parameters, such as atmosphere, duration and catalyst bed depth have a marked influence on the catalytic and spectroscopic properties of sulfated zirconia. Sulfated zirconia calcined in nitrogen or synthetic airflow, in deep bed, exhibited comparable activity in n-butane isomerization at 373 K, which suggests that oxygen is not necessary for formation of active sites. Catalysts calcined in shallow bed are catalytically inactive. Thus, the bed depth is concluded to be crucial for the formation of active sites. The samples calcined in shallow bed possessed lower sulfate content and the S=O stretching vibration was located at lower frequency. Calcination in the presence of water vapor also led to lower catalytic activity, sulfate content, and BET area. Extended calcination reduced gradually the activity and the sulfate content, which underlines the labile property of the active sites. A new interpretation of the function of the calcination step is proposed and compared with models described in the literature.
Co-reporter:Stefan Gaab;Josef Find;Thomas E. Müller
Topics in Catalysis 2007 Volume 46( Issue 1-2) pp:101-110
Publication Date(Web):2007 September
DOI:10.1007/s11244-007-0320-x
The microkinetic reaction network of the oxidative dehydrogenation of ethane to ethene over Li/Dy/Mg/O and Li/Dy/Mg/O/Cl catalysts was investigated. With Li/Dy/Mg/O catalysts, the reaction kinetics is compatible with a heterogeneous-homogeneous radical based reaction mechanism. The formation of ethyl radicals on the surface is concluded to be the rate-determining step. In contrast, the reaction kinetics for Li/Dy/Mg/O/Cl is in line with a purely surface catalyzed reaction mechanism. However, also in this case, alkane activation is rate determining.
Co-reporter:Hendrik Dathe, Andreas Jentys, Peter Haider, Ellen Schreier, Rolf Fricke and Johannes A. Lercher  
Physical Chemistry Chemical Physics 2006 vol. 8(Issue 13) pp:1601-1613
Publication Date(Web):18 Jan 2006
DOI:10.1039/B515678E
Calcium–aluminum mixed oxide based materials doped with Na and Mn were explored as sulfur trapping materials. The materials showed a three times higher total storage capacity and a higher time on stream with complete SO2 removal compared to a second generation SOx trapping material which was mesoporous with calcium mainly present in oxidic form. Combining in situ XANES at the S K-edge and IR spectroscopy the key properties of the storage materials and the affiliated storage processes were identified. CaO–Al2O3 acts as the primary support and storage component, while Na+ cations adjust the base strength and enhances the storage capacity. Manganese cations provide the appropriate oxidation capacity in absence and presence of up to 10% water. The transport into the bulk phase, which is markedly influenced by a layer of sorbed water, is the rate-limiting step in presence of Mn cations. In the absence of manganese cations the oxidation step appears controlling the rate. The overall reaction network, identified by in situ IR spectroscopy and the 2D Correlation Analysis, is similar on all materials.
Co-reporter:Hendrik Dathe, Andreas Jentys and Johannes A. Lercher  
Physical Chemistry Chemical Physics 2005 vol. 7(Issue 6) pp:1283-1292
Publication Date(Web):18 Feb 2005
DOI:10.1039/B419077G
The elementary steps during oxidative chemisorption of SO2 by a novel composite material consisting of highly disordered benzene tri-carboxylate metal organic framework materials with Cu as central cation and BaCl2 as a second component (Ba/Cu-BTC) and by a conventional BaCO3/Al2O3/Pt based material were investigated. EXAFS analysis on the Cu K-edge in Ba/Cu-BTC indicates the opening of the majority of the Cu–Cu pairs present in the parent Cu-BTC. Compared to Cu-BTC, the BaCl2 loaded material has hardly any micropores and has higher disorder, but it has better accessibility of the Cu2+ cations. This results from the partial destruction of the MOF structure by reaction between BaCl2 and the Cu cations. The SO2 uptake in oxidative atmosphere was higher for the Ba/Cu-BTC sample than for the BaCO3/Al2O3/Pt based material. XRD showed that on Ba/Cu-BTC the formation of BaSO4 and CuSO4 occurs in parallel to the destruction of the crystalline structure. With BaCO3/Al2O3/Pt the disappearance of carbonates was accompanied with the formation of Ba- and Al-sulfates. XANES at the S K-edge was used to determine the oxidation states of sulfur and to differentiate between the sulfate species formed. At low temperatures (473 K) BaSO4 was formed preferentially (53 mol% BaSO4, 47 mol% CuSO4), while at higher temperatures (and higher sulfate loading) CuSO4 was the most abundant species (42 mol% BaSO4, 58 mol% CuSO4). In contrast, on the BaCO3/Al2O3/Pt based material the relative concentration of the sulfate species (i.e., BaSO4 and Al2(SO4)3) as function of the temperature remained constant.
Co-reporter:J. Lerchre
Chemie Ingenieur Technik 2005 Volume 77(Issue 8) pp:
Publication Date(Web):10 AUG 2005
DOI:10.1002/cite.200590269
Co-reporter:Christian Sedlmair, Barbara Gil, Kulathuiyer Seshan, Andreas Jentys and Johannes A. Lercher  
Physical Chemistry Chemical Physics 2003 vol. 5(Issue 9) pp:1897-1905
Publication Date(Web):25 Mar 2003
DOI:10.1039/B209325A
The surface species formed during adsorption of NO, NO+O2 and NO2 on sodium and barium exchanged Y zeolites have been investigated by in situ IR spectroscopy. Ionically bound nitrates and nitrites on the exchanged metal cation sites are the main species formed during adsorption of NO and NO2. Extra framework alumina was identified as additional sorption site forming small concentrations of bridging, chelating and monodentate nitrates. N2O4 and NO+ were found to be reaction intermediates during the NOx adsorption process. The direct oxidation of NO2 with reactive oxygen from the zeolite surface is facilitated by the formation of nitrates via the disproportionation reaction of N2O4 to NO+ and NO3−. NO+ was found to act as precursor for the creation of nitrites. Decomposition of the nitrate species occurs between 150 and 450°C. During the temperature increase less stable nitrite/nitrate species are transformed into Ba-nitrates showing the highest thermal stability. The stability of surface nitrates/nitrites was found to be lower, if NO instead of NO2 is present in the feed during temperature increase. For the interaction of surface NOx species with propene two pathways are proposed. At low temperatures, NO+ was identified as the active NOx surface species reacting with propene to nitriles. At higher temperatures the reduction of surface nitrates/nitrites occurred via organic nitro/nitrito species, carboxylic species and isocyanates.
Co-reporter:R.Q. Su;T.E. Müller;J. Procházka;J.A. Lercher
Advanced Materials 2002 Volume 14(Issue 19) pp:
Publication Date(Web):27 SEP 2002
DOI:10.1002/1521-4095(20021002)14:19<1369::AID-ADMA1369>3.0.CO;2-I
Co-reporter:Jan-Olaf Barth, Jan Kornatowski and Johannes A. Lercher  
Journal of Materials Chemistry A 2002 vol. 12(Issue 2) pp:369-373
Publication Date(Web):03 Jan 2002
DOI:10.1039/B104824B
Alumina and magnesia–alumina have been applied for pillaring the layered precursor of zeolite MCM-22. Calcination of such materials has led to new varieties of MCM-36 molecular sieves. Pillaring with alumina yields mesoporous materials with lower surface areas than those pillared with silica. The pillaring process with alumina strongly depends upon the preparation conditions of the alumina species and requires an elongated aging of the pillaring solutions. Application of magnesia in addition to alumina results in a higher exfoliation of the MCM-22 layers and incorporation of an increased amount of alumina into the materials. MgO × Al2O3 as pillaring agent yields a significantly higher mesoporosity in the range of 2–4 nm as compared with Al2O3. This synthesis method is a promising tool for tailoring the physicochemical properties and the structure of the slit-mesopores in relation to materials typically pillared with silica.
Co-reporter:Gautam S Nivarthy, Andreas Feller, Kulathuiyer Seshan, Johannes A Lercher
Microporous and Mesoporous Materials 2000 Volumes 35–36() pp:75-87
Publication Date(Web):April 2000
DOI:10.1016/S1387-1811(99)00209-7
Alkylation of isobutane with ethene and propene was studied over an H-BEA catalyst in a well-stirred reactor. Under similar conditions of space velocity and paraffin-to-olefin feed ratio, lower initial olefin conversions were observed with ethene or propene than those reported earlier for butene. The sole presence of saturated hydrocarbons indicates the dominance of the alkylation over alkene oligomerization. The product distribution indicates that the rate of hydride transfer in the reaction sequence of alkylation is significantly lower than the rate of methyl shift isomerization. For ethene, significant amounts of C8 alkanes, especially trimethylpentanes were found in the products indicating multiple alkylation or dimerization followed by alkylation to dominate. With propene, iso-heptanes dominated the products, as expected for single alkylation. Next to single and multiple alkylation, evidence for cracking of surface alkoxy-species by β-scission, iso-butane self-alkylation and olefin dimerization was observed. The product distribution suggests that dimerization over weaker acidic sites allowing desorption of olefins occurred also during the period of stable catalytic activity, but the resulting olefins were immediately added to alkoxy species present and thus contributed to the alkylation route. Once these alkoxy species become large, either olefin addition or hydride transfer is blocked, and olefins are detected in the products.
Co-reporter:L Domokos, L Lefferts, K Seshan, J.A Lercher
Journal of Molecular Catalysis A: Chemical 2000 Volume 162(1–2) pp:147-157
Publication Date(Web):20 November 2000
DOI:10.1016/S1381-1169(00)00286-7
The role of the acid site location in H-ferrierite (H-FER) on the skeletal isomerization of linear butenes was studied. Assignment of OH groups observed in FTIR analysis is addressed taking into account previous NMR and computational studies regarding the FER structure. Possible locations of alkali ions in the zeolite framework are discussed. Close structure activity relationship has been observed between Brønsted acid sites located in the 10 member ring (MR) channels and selective isobutene formation. Molecular dynamics (MD) calculations of possible products observed during n-butene transformation at reaction temperatures support the experimental findings.
Co-reporter:Oliver Y. Gutiérrez, Ana Hrabar, Jennifer Hein, Yanzhe Yu, Jinyi Han, Johannes A. Lercher
Journal of Catalysis (November 2012) Volume 295() pp:155-168
Publication Date(Web):1 November 2012
DOI:10.1016/j.jcat.2012.08.003
The hydrodenitrogenation of decahydroquinoline (DHQ) and quinoline on MoS2/γ-Al2O3 and Ni–MoS2/γ-Al2O3 proceeds via two routes. The first one proceeds via DHQ → propylcyclohexylamine → propylcyclohexene → propylcyclohexane, and the ring opening in DHQ is the rate-limiting step. The second route proceeds via 1,2,3,4-tetrahydroquinoline (14THQ) → o-propylaniline → propylcyclohexylamine and propylbenzene with the ring opening of 14THQ and the hydrogenation of o-propylaniline being the rate determining steps (the intrinsic rate of C(sp3)–N bond cleavage being slower in 14THQ than in DHQ). The active sites for the ring opening via Hofmann elimination are acidic –SH groups and basic S2− ions. The parallel conversion of dibenzothiophene (DBT) via direct desulfurization provides increasing concentrations of S2− ions and –SH groups. Nickel facilitates the adsorption of H2S and H2 and the mobility of hydrogen. Thus, the presence of DBT and Ni accelerates the rate of the C(sp3)–N bond cleavage. H2S as sulfur source enhances the ring-opening steps in a minor extent than DBT. The presence of –SH groups and the effect of Ni on them were probed by TPR, TPD and IR-spectroscopy of adsorbed 2,6-dimethylpyridine.Graphical abstractThe ring opening of hydrogenated intermediates via Hofmann elimination is the limiting step in the hydrodenitrogenation of quinoline. The direct desulfurization of dibenzothiophene and the presence of Ni facilitate the formation of acidic –SH groups and basic S2− ions, therefore, accelerate the rate of the C(sp3)–N bond cleavage.Download high-res image (129KB)Download full-size imageHighlights► Hydrodenitrogenation of decahydroquinoline and quinoline on (Ni)MoS2/γ-Al2O3. ► The C–N cleavage in tetrahydroquinoline and decahydroquinoline is the determining step. ► Dibenzothiophene and Ni accelerate the rate of that C(sp3)–N bond cleavage. ► The ring opening via Hofmann elimination occurs on acidic –SH groups and S2− ions.
Co-reporter:Hui Shi, Xuebing Li, Gary L. Haller, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (November 2012) Volume 295() pp:133-145
Publication Date(Web):1 November 2012
DOI:10.1016/j.jcat.2012.08.005
Hydrogenolysis of cyclohexane has been explored over supported Ir catalysts. The kinetic data are combined with modeling results to assess the structural requirements and the nature of catalytically relevant surface intermediates for endocyclic C–C bond cleavage. The turnover frequency (TOF) for cyclohexane hydrogenolysis showed complex dependence on Ir particle size, while the selectivity to the primary ring opening product, n-hexane, decreased monotonically with decreasing Ir dispersion. The decreasing TOF as the Ir dispersion decreased from 65% to 52% originates principally from the diminishing abundance of low-coordination Ir atoms at particle surfaces. The increase of the TOF with further Ir particle growth is attributed to an increased fraction of terrace planes, or step sites, and a less unsaturated nature of the most abundant reactive intermediate. Selectivities for multiple C–C bond cleavage, yielding C<6 alkanes, varies with the relative abundance of coordinatively unsaturated Ir atoms and terrace planes. The multiple hydrogenolysis depends additionally upon H2 pressure, because single and multiple C–C bond scissions are mediated by surface intermediates with different H-deficiencies.Graphical abstractThe turnover frequency (TOF) for cyclohexane hydrogenolysis on supported Ir catalysts decreases as the dispersion decreases from 65% to 52% due to the diminishing abundance of low-coordination Ir atoms. The TOF increases with further Ir particle growth due to a varying nature of the most abundant reactive intermediate on terraces.Download high-res image (123KB)Download full-size imageHighlights► Structure sensitivity of cyclohexane ring opening (RO) is studied on Ir catalysts. ► The variation of TOF values presents a minimum at intermediate Ir dispersions. ► The TOF change on small particles is related to the fraction of low-CN atoms. ► Changed nature and size requirement of MARI cause TOF to increase on large particles. ► The relevant MARI is more H-depleted for the multiple hydrogenolysis than for RO.
Co-reporter:Xianyong Sun, Sebastian Mueller, Hui Shi, Gary L. Haller, Maricruz Sanchez-Sanchez, Andre C. van Veen, Johannes A. Lercher
Journal of Catalysis (May 2014) Volume 314() pp:21-31
Publication Date(Web):1 May 2014
DOI:10.1016/j.jcat.2014.03.013
•Effect of feed composition was studied under reaction conditions close to industrial MTP process.•Co-feeding aromatics with methanol enhanced the aromatic-based catalytic cycle.•Co-feeding of C3–C6 olefins did not change selectivity of ZSM-5 toward products.•A way to increase propene yield and tune the ethene-to-propylene ratio is presented.The impact of adding various aromatic molecules (benzene, toluene, and xylenes) or olefins (ethene, propene, 1-butene, 1-pentene, and 1-hexene) to methanol over a HZSM-5 catalyst on activity and selectivity was systematically studied. Addition of a low concentration of aromatic molecules (16–32 C%), which are free of diffusion constraints, significantly enhanced the aromatics-based catalytic cycle and greatly suppressed the olefin-based cycle. This led to enhanced methane and ethene formation and methylation of aromatic rings at the expense of propene and C4+ higher olefins. The ratio of propene to ethene is controlled by the concentration of the aromatic molecules added. Co-feeding the same molar concentration of benzene, toluene and p-xylene influenced the methanol conversion to a nearly identical extent, as none of them experience transport constraints and the methylation rapidly equilibrates the aromatic molecules retained in the pores. In stark contrast, addition of small concentrations (10–40 C%) of C3–6 olefins with 100 C% methanol does not selectively suppress the catalytic cycle based on aromatic molecules. This led to unchanged selectivities to ethene and higher olefins (C3+). Within the C3+ fraction, the selectivity to propene decreased and the selectivity to butenes were enhanced with increasing concentration of the co-fed olefin. Because of the relatively fast rates in methylation and cracking of C3–6 olefins in the olefin-based cycle, the product distributions at high methanol conversion were identical when co-feeding C3–6 olefins with the same carbon concentrations. This work provides further insights into the two distinct catalytic cycles operating for the methanol conversion to produce ethene and propene over HZSM-5 catalysts.Download high-res image (108KB)Download full-size image
Co-reporter:Yang Song, Oliver Y. Gutiérrez, Juan Herranz, Johannes A. Lercher
Applied Catalysis B: Environmental (March 2016) Volume 182() pp:
Publication Date(Web):1 March 2016
DOI:10.1016/j.apcatb.2015.09.027
•Efficient hydrogenation of phenol at room temperature and atmospheric pressure.•Electrochemical and thermal hydrogenations on Pt/C and Rh/C are compared.•Both routes are independent but have the rate same rate determining step (RDS).•The RDS is the hydrogenation of adsorbed hydrocarbons.The electrocatalytic hydrogenation (ECH) of phenol on Pt/C, Rh/C, and Pd/C was explored in an H-type two-compartment cell with respect to the impact of electrolyte, pH, current, and catalyst concentration. In all cases, the electric efficiencies increased with increasing phenol conversions. Rh/C exhibited the highest hydrogenation rate normalized to the concentration of accessible metal (TOF) followed by Pt/C in terms of mass of metal and intrinsic activities. Therefore, the effect of temperature on ECH and of mild thermal hydrogenation (TH) of phenol was explored on these catalysts. The activation energies for ECH were ca. 23 kJ/mol and 29 kJ/mol on Rh/C, and Pt/C, respectively. TH is much faster than ECH, although both pathways have the same activation energy. Cyclic voltammetry of bulk Pt and Pt/C in the presence of phenol indicated that phenol is adsorbed on the metal and reacted with hydrogen radicals. Hence, ECH was concluded to proceed via a Langmuir-type mechanism where the surface hydrogen is produced by reduction of protons (which occurs when the catalyst contacts the electrode) instead of H2 dissociation as in TH. Although competitive reactions evolve H2 during ECH, the involvement of this H2 in phenol hydrogenation was minor. Thus, ECH and TH are independent processes and do not exhibit any synergy. In both pathways, the reaction path is phenol → cyclohexanone → cyclohexanol. CO bond cleavage was not observed.Download high-res image (90KB)Download full-size image
Co-reporter:Jiayue He, Lu Lu, Chen Zhao, Donghai Mei, Johannes A. Lercher
Journal of Catalysis (March 2014) Volume 311() pp:41-51
Publication Date(Web):1 March 2014
DOI:10.1016/j.jcat.2013.10.024
•Benzyl phenyl ether is cleaved by hydrolysis in water with hydronium ions from water or solid acids at 523 K.•Ni/SiO2 catalyzes the dominant hydrogenolysis of the C−O bond of benzyl phenyl ether in water.•Hydrogenolysis occurs as the major route rather than hydrolysis on Ni/HZSM-5 in water.•In undecane, non-catalytic thermal pyrolysis route dominates on benzyl phenyl ether conversion.•Hydrogenolysis appears to be the major reaction pathway in undecane with metal sites.Catalytic pathways for the cleavage of ether bonds in benzyl phenyl ether (BPE) in liquid phase using Ni- and zeolite-based catalysts are explored. In the absence of catalysts, the C−O bond is selectively cleaved in water by hydrolysis, forming phenol and benzyl alcohol as intermediates, followed by alkylation. The hydronium ions catalyzing the reactions are provided by the dissociation of water at 523 K. Upon addition of HZSM-5, rates of hydrolysis and alkylation are markedly increased in relation to proton concentrations. In the presence of Ni/SiO2, the selective hydrogenolysis dominates for cleaving the Caliphatic−O bond. Catalyzed by the dual-functional Ni/HZSM-5, hydrogenolysis occurs as the major route rather than hydrolysis (minor route). In apolar undecane, the non-catalytic thermal pyrolysis route dominates. Hydrogenolysis of BPE appears to be the major reaction pathway in undecane in the presence of Ni/SiO2 or Ni/HZSM-5, almost completely suppressing radical reactions. Density functional theory (DFT) calculations strongly support the proposed C−O bond cleavage mechanisms on BPE in aqueous and apolar phases. These calculations show that BPE is initially protonated and subsequently hydrolyzed in the aqueous phase. DFT calculations suggest that the radical reactions in non-polar solvents lead to primary benzyl and phenoxy radicals in undecane, which leads to heavier condensation products as long as metals are absent for providing dissociated hydrogen.Graphical abstractDownload high-res image (184KB)Download full-size image
Co-reporter:John H. Ahn, Robin Kolvenbach, Carolina Neudeck, Sulaiman S. Al-Khattaf, Andreas Jentys, Johannes A. Lercher
Journal of Catalysis (March 2014) Volume 311() pp:271-280
Publication Date(Web):1 March 2014
DOI:10.1016/j.jcat.2013.12.003
•Design of catalysts for toluene methylation based on mesoscopically structured H-ZSM5.•Path length reduction by post-synthetic treatments is beneficial for catalyst activity.•SiO2 overlayer increases p-xylene selectivity due to enhanced tortuosity.•Final catalyst material combines better utilization with higher p-xylene selectivity.•Exploring influence of transport and acidic properties on catalyst performance.Mesoscopically structured zeolites based on H-ZSM5 were designed and synthesized as highly active and shape selective catalysts for methylation of toluene by tuning diffusion and acid site concentration of the catalysts. This was achieved by combining desilication, subsequent dealumination and chemical deposition of a mesoporous SiO2 overlayer of several nanometer thickness. The decreasing effective diffusion length in zeolite crystals achieved by desilication and dealumination increased the turnover rate of toluene by favoring activation of methanol and facilitating desorption of the produced xylenes, albeit with some loss in p-xylene selectivity. The presence of the SiO2 overlayer increased the p-xylene selectivity by enhancing the tortuosity of the zeolite, randomly blocking pore openings at the surface, and increasing the effective diffusion path length. The final material combines the higher catalyst utilization with enhanced selectivity leading to rates comparable to the parent zeolite, but at significantly higher selectivity.Graphical abstractDownload high-res image (125KB)Download full-size image
Co-reporter:S. Schallmoser, T. Ikuno, M.F. Wagenhofer, R. Kolvenbach, G.L. Haller, M. Sanchez-Sanchez, J.A. Lercher
Journal of Catalysis (July 2014) Volume 316() pp:93-102
Publication Date(Web):1 July 2014
DOI:10.1016/j.jcat.2014.05.004
•n-Pentane cracking studied over MFI in a wide range of acid site concentrations.•Tetrahedrally coordinated EFAl in proximity to BAS is responsible for higher rates.•Equal turnover frequency per BAS was found for samples free of EFAl.•Post-synthetic chemical modification enables tuning of catalytic activity.The impact of the zeolite Brønsted and Lewis acid site concentration on the catalytic cracking of alkanes was explored using n-pentane and H-ZSM-5 as examples. Rates normalized to strong Brønsted acid sites (i.e., the turnover frequencies, TOF) showed that the two samples with the highest Al content had much higher TOF than all other samples. This difference has been unequivocally linked to the presence of extra-lattice alumina. Post-treatment of the zeolites with ammonium hexafluorosilicate and static calcination was used to vary the concentration of extra-lattice alumina. After extraction of extra-lattice alumina from the samples with high TOF, all TOFs were identical. IR spectra of adsorbed pyridine and NH3, coupled with 27Al MAS NMR, showed that the overall enhanced activity is associated with tetrahedrally coordinated extra-lattice alumina in close proximity to strong Brønsted acid sites. The TOF of these sites is approximately 40-times higher than the TOF on normal Brønsted acid sites.Graphical abstractDownload high-res image (133KB)Download full-size image
Co-reporter:Anastasia V. Pashigreva, Elena Kondratieva, Ricardo Bermejo-Deval, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (January 2017) Volume 345() pp:308-318
Publication Date(Web):1 January 2017
DOI:10.1016/j.jcat.2016.11.036
•Reactions of methanol and H2S over Al2O3 and WS2/Al2O3 with or without Cs+.•Dimethyl ether and methanethiol form on strong acid sites of the Cs+-free catalysts.•Reactions follow bimolecular Langmuir-Hinshelwood (SN2) mechanisms.•Strong acid sites of the parent materials are replaced by Cs+ with weak acidity.•Cs+ sites only catalyze methanol thiolation to methanethiol suppressing side reactions.Thiolation of methanol with H2S was studied on Al2O3, WS2/Al2O3, and the Cs-modified counterparts Cs/Al2O3, and Cs-WS2/Al2O3. On Cs-free catalysts, methanol reacts with similar rates to dimethyl ether and to methanethiol. Secondary steps yield dimethyl sulfide and trace amounts of dimethyl disulfide and methane. On Cs-containing catalysts, the dominating reaction is methanol thiolation. This drastic change in selectivity is related to the low strength and concentration of Lewis acid sites in the presence of Cs+. The active sites for condensation and thiolation of methanol are concluded to be Lewis or Brønsted acid-base pairs formed by dissociation of H2S on the Lewis sites. Thiolation and condensation follow Langmuir-Hinshelwood mechanisms with nucleophilic attack of SH or methoxy groups to adsorbed methanol as rate determining step. The formation of dimethyl ether on Cs+ Lewis acid sites is suppressed due to low coverage of methanol, whereas the rate of the nucleophilic attack of SH to methanethiol is enhanced.Download high-res image (148KB)Download full-size image
Co-reporter:Jiayue He, Chen Zhao, Donghai Mei, Johannes A. Lercher
Journal of Catalysis (January 2014) Volume 309() pp:280-290
Publication Date(Web):1 January 2014
DOI:10.1016/j.jcat.2013.09.012
•Diphenyl ether is cleaved by parallel hydrogenolysis and hydrolysis on Ni.•Hydrogenolysis is the exclusive route for cleaving a C–O bond of di-p-tolyl ether with Ni.•H* compete with reactant for adsorption as a function of H2 pressure.•4,4′-Dihydroxydiphenyl ether undergoes sequential surface hydrogenolysis with Ni.•The rate order: 4,4′-dihydroxydiphenyl ether > diphenyl ether > di-p-tolyl ether over Ni.A route for cleaving the C–O aryl ether bonds of p-substituted H–, CH3–, and OH– diphenyl ethers has been explored over Ni/SiO2 catalyst at very mild conditions (393 K, 0.6 MPa). The C–O bond of diphenyl ether is cleaved by parallel hydrogenolysis and hydrolysis (hydrogenolysis combined with HO* addition) on Ni. The rates as a function of H2 pressure from 0 to 10 MPa indicate that the rate-determining step is the C–O bond cleavage on Ni surface. H* atoms compete with the organic reactant for adsorption leading to a maximum in the rate with increasing H2 pressure. In contrast to diphenyl ether, hydrogenolysis is the exclusive route for cleaving a C–O bond of di-p-tolyl ether to form p-cresol and toluene. 4,4′-Dihydroxydiphenyl ether undergoes sequential surface hydrogenolysis, first to phenol and OC6H4OH* (adsorbed), which is then cleaved to phenol (C6H4OH* with added H*) and H2O (O* with two added H*) in a second step. Density function theory supports the operation of this pathway. Notably, addition of H* to OC6H4OH* is less favorable than a further hydrogenolytic C–O bond cleavage. The TOFs of three diaryl ethers with Ni/SiO2 in water follow the order 4,4′-dihydroxydiphenyl ether 69molmolNi Surf-1h-1 > diphenyl ether 26molmolNi Surf-1h-1 > di-p-tolyl ether 1.3molmolNi Surf-1h-1, in line with the increasing apparent activation energies, ranging from 4,4′-dihydroxydiphenyl ether (93 kJ mol−1) < diphenyl ether (98 kJ mol−1) < di-p-tolyl ether (105 kJ mol−1).Graphical abstractDownload high-res image (199KB)Download full-size image
Co-reporter:Jiayue He, Chen Zhao, Johannes A. Lercher
Journal of Catalysis (January 2014) Volume 309() pp:362-375
Publication Date(Web):1 January 2014
DOI:10.1016/j.jcat.2013.09.009
•Solvents impact reductive deoxygenation of phenol by altering H2, substrate, and intermediate interactions.•Methanol provides an additional acetal reaction route of cyclohexanone with methanol.•Methanol not only strongly solvates phenol and cyclohexanone, but also adsorbs strongly on Pd.•The low solubility of phenol and strong interaction of hexadecane–Pd surface in hexadecane lead to slower rates.Impacts of water, methanol, and hexadecane solvents on the individual steps of phenol hydrodeoxygenation are investigated over Pd/C and HZSM-5 catalyst components at 473 K in presence of H2. Hydrodeoxygenation of phenol to cyclohexane includes four individual steps of phenol hydrogenation to cyclohexanone on Pd/C, cyclohexanone hydrogenation to cyclohexanol on Pd/C, cyclohexanol dehydration to cyclohexene on HZSM-5, and cyclohexene hydrogenation to cyclohexane on Pd/C. Individual phenol and cyclohexanone hydrogenation rates are much lower in methanol and hexadecane than in water, while rates of cyclohexanol dehydration and cyclohexene hydrogenation are similar in three solvents. The slow rate in methanol is due to the strong solvation of reactants and the adsorption of methanol on Pd, as well as to the reaction between methanol and the cyclohexanone intermediate. The low solubility of phenol and strong interaction of hexadecane with Pd lead to the slow rate in hexadecane. The apparent activation energies for hydrogenation follow the order Ea phenol > Ea cyclohexanone > Ea cyclohexene, and the sequences of individual reaction rates are reverse in three solvents. The dehydration rates 1.1–1.8×103molmolBAS-1h-1 and apparent activation energies (115–124 kJ mol−1) are comparable in three solvents. In situ liquid-phase IR spectroscopy shows the rates consistent with kinetics derived from chromatographic evidence in the aqueous phase and verifies that hydrogenation of phenol and cyclohexanone follows reaction orders of 1.0 and 0.55 over Pd/C, respectively. Conversion of cyclohexanol with HZSM-5 shows first-order dependence in approaching the dehydration–hydration equilibrium in the aqueous phase.Download high-res image (160KB)Download full-size image
Co-reporter:Hui Shi, Oliver Y. Gutiérrez, Gary L. Haller, Donghai Mei, Roger Rousseau, Johannes A. Lercher
Journal of Catalysis (January 2013) Volume 297() pp:70-78
Publication Date(Web):1 January 2013
DOI:10.1016/j.jcat.2012.09.018
Structure sensitivities, H2 pressure effects, and temperature dependencies for rates and selectivities of endo- and exocyclic C–C bond cleavage in methylcyclohexane were studied over supported Ir catalysts. The rate of endocyclic C–C bond cleavage first decreased and then increased with declining Ir dispersion from 0.65 to 0.035. The ring opening (RO) product distribution remained unchanged with varying H2 pressure on small Ir particles, while further shifting to methylhexanes with increasing H2 pressure on large particles. In contrast, the rate and selectivity of exocyclic C–C bond cleavage decreased monotonically with increasing H2 pressure and decreasing Ir particle size. The distinct dependencies of endocyclic and exocyclic C–C bond cleavage pathways on Ir dispersion and H2 pressure suggest that they are mediated by surface species with different ensemble size requirements. DFT calculations were performed on an Ir50 cluster and an Ir(1 1 1) surface, with or without pre-adsorbed hydrogen atoms, to provide insight into the observed effects of particle size and H2 pressure on RO pathways. On small Ir particles, the calculated dehydrogenation enthalpies for all endocyclic bonds were similar and affected to similar extents by H2 pressure; on large particles, the selectivity to n-heptane (via substituted C–C bond cleavage) was even lower than on small particles as a result of the least favorable adsorption and dehydrogenation energetics for hindered bonds.Graphical abstractThe turnover frequency (TOF) for methylcyclohexane hydrogenolysis on supported Ir catalysts decreases as the dispersion decreases from 65% to 46%. The TOF increases with further Ir particle growth. The selectivity along the demethylation pathway, while being negligible on corners and edges, becomes comparable to ring opening on terrace planes.Download high-res image (54KB)Download full-size imageHighlights► Rates of endocyclic C–C cleavage feature a minimum at medium Ir dispersion. ► TOFs of demethylation monotonously increase with increasing Ir particle size. ► Ring opening (RO) selectivity decreases by up to three times with declining dispersion. ► Endocyclic and exocyclic cleavage proceed via different surface intermediates. ► DFT calculations show unfavored adsorption/dehydrogenation energetics at hindered C–C.
Co-reporter:Kai E. Sanwald, Tobias F. Berto, Wolfgang Eisenreich, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (December 2016) Volume 344() pp:806-816
Publication Date(Web):1 December 2016
DOI:10.1016/j.jcat.2016.08.009
•Photoreforming and oxidation mechanisms of polyols on Rh/TiO2 for H2-generation.•Anodic reaction network for glycerol photoreforming was quantitatively elucidated.•Oxidative C–C cleavage is attributed to direct transfer of photogenerated holes.•Formation of carbonyl moieties and dehydration result from indirect hole transfer.•C–C bond rupture is favored with increasing polyol carbon number.Photocatalytic reforming of biomass-derived oxygenates leads to H2 generation and evolution of CO2 via parallel formation of organic intermediates through anodic oxidations on a Rh/TiO2 photocatalyst. The reaction pathways and kinetics in the photoreforming of C3–C6 polyols were explored. Polyols are converted via direct and indirect hole transfer pathways resulting in (i) oxidative rupture of C–C bonds, (ii) oxidation to α-oxygen functionalized aldoses and ketoses (carbonyl group formation) and (iii) light-driven dehydration. Direct hole transfer to chemisorbed oxygenates on terminal Ti(IV)-OH groups, generating alkoxy-radicals that undergo ß-C–C-cleavage, is proposed for the oxidative C–C rupture. Carbonyl group formation and dehydration are attributed to indirect hole transfer at surface lattice oxygen sites [Ti⋯O⋯Ti] followed by the generation of carbon-centered radicals. Polyol chain length impacts the contribution of the oxidation mechanisms favoring the C–C bond cleavage (internal preferred over terminal) as the dominant pathway with higher polyol carbon number.Download high-res image (73KB)Download full-size image
Co-reporter:Yang Song, Shao Hua Chia, Udishnu Sanyal, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (December 2016) Volume 344() pp:263-272
Publication Date(Web):1 December 2016
DOI:10.1016/j.jcat.2016.09.030
•Electrocatalytic hydrogenation of phenolic compounds and diaryl ethers on Rh/C.•The reaction proceeds at room temperature without external H2.•Hydrogenation of the aromatic rings is the dominant reaction pathway.•Methoxy and benzyloxy groups undergo C–O bond cleavage at room temperature.•Increasing cathodic potentials favors the rates of ECH over H2 evolution.Electrocatalytic hydrogenation and catalytic thermal hydrogenation of substituted phenols and diaryl ethers were studied on carbon-supported Rh. The rates of electrocatalytic hydrogenation increase with increasingly negative potentials, which have been related with the coverage of adsorbed hydrogen. For electrocatalytic and catalytic thermal hydrogen addition reactions, the dominant reaction pathway is hydrogenation to cyclic alcohols and cycloalkyl ethers. The presence of substituting methyl or methoxy groups led to lower rates compared to unsubstituted phenol or diphenyl ether. Methoxy or benzyloxy groups, however, undergo C–O bond cleavage via hydrogenolysis and hydrolysis (minor pathway). The surface chemical potential of hydrogen can be increased also by generating a H2 atmosphere above the reaction media, supporting the conclusion that thermal and electrochemical routes share the same reaction pathways.Download high-res image (130KB)Download full-size image
Co-reporter:Chen Zhao, Stanislav Kasakov, Jiayue He, Johannes A. Lercher
Journal of Catalysis (December 2012) Volume 296() pp:12-23
Publication Date(Web):1 December 2012
DOI:10.1016/j.jcat.2012.08.017
We have investigated the detailed kinetics of phenol hydrodeoxygenation in liquid aqueous medium over Ni supported on HZSM-5 or HZSM-5 with 19.3 wt.% γ-Al2O3 binder. The individual reaction steps on two Ni catalysts followed the rate order r1 (phenol hydrogenation) < r2 (cyclohexanone hydrogenation) < r3 (cyclohexanol dehydration) ≪ r4 (cyclohexene hydrogenation), so that phenol hydrogenation was the rate determining step. As Ni/Al2O3-HZSM-5 showed up to five times higher catalytic activity for phenol hydrogenation than Ni/HZSM-5, it also delivered higher rates for overall phenol hydrodeoxygenation, verified by both the kinetic study monitored by GC and in situ IR spectroscopy to trace the product concentrations. Under the present conditions, Ni leaching was almost negligible from Ni/Al2O3-HZSM-5 after 90 h. The HZSM-5 support was stable, but the Al2O3-HZSM-5 support lost 7 wt.%. The catalytic activity gradually decreased when catalyst was recovered and reused, mainly due to Ni particle sintering tracked by transmission electron microscopy. The change of acid sites on the fresh and used catalysts monitored by IR of adsorbed pyridine demonstrated that Brønsted acid sites of HZSM-5 could be irreversibly transformed to Lewis acid sites during calcination.Graphical abstractDownload high-res image (144KB)Download full-size imageHighlights► The rates of individual reactions of phenol hydrodeoxygenation were compared on two Ni catalysts. ► Ni/Al2O3-HZSM-5 was more active due to higher Ni dispersion. ► The sequential rates r1 < r2 < r3 << r4 in phenol hydrodeoxygenation. ► The catalysts deactivated due to Ni sintering during the hydrothermal and recycling process. ► By calcination BAS were dehydrated to coordinately unsaturated Al Lewis acid sites.
Co-reporter:Xianyong Sun, Sebastian Mueller, Yue Liu, Hui Shi, Gary L. Haller, Maricruz Sanchez-Sanchez, Andre C. van Veen, Johannes A. Lercher
Journal of Catalysis (August 2014) Volume 317() pp:185-197
Publication Date(Web):1 August 2014
DOI:10.1016/j.jcat.2014.06.017
•MTO mechanism was explored under industrially relevant conditions on ZSM-5.•Olefin- and aromatic-based cycle contribution depends on local chemical potential.•Co-feeding small concentrations of higher alcohols significantly increased catalyst lifetime.•Hydride transfer involving methanol appears to be responsible for a marked decrease in activity.The underlying mechanisms of the two distinct catalytic cycles operating during conversion of methanol to olefins (MTO) on HZSM-5 have been elucidated under industrially relevant conditions. The co-existence of olefins and aromatic molecules in the zeolite pores leads to competition between the two cycles. Therefore, their importance depends on the local chemical potential of specific carbon species and the methanol conversion. Due to a faster, “autocatalytic” reaction pathway in the olefin based cycle, olefin homologation/cracking is dominant under MTO conditions, irrespective of whether aromatic molecules or olefins are co-fed with methanol. Another hydrogen transfer pathway, faster than the usual route, has been identified, which is directly linked to methanol. In agreement with that, the co-feeding of olefins resulted in a remarkable longer lifetime of the catalyst under MTO conditions, because the high rate methylation competes with the formation of more deactivating coke – presumably oxygenates- through methanol derivatives.Graphical abstractDownload high-res image (148KB)Download full-size image
Co-reporter:Yanzhe Yu, Oliver Y. Gutiérrez, Gary L. Haller, Robert Colby, Bernd Kabius, J.A. Rob van Veen, Andreas Jentys, Johannes A. Lercher
Journal of Catalysis (August 2013) Volume 304() pp:135-148
Publication Date(Web):1 August 2013
DOI:10.1016/j.jcat.2013.04.009
•Synthesis and characterization of ASA-supported Pt, Pd, and Pt–Pd catalysts.•Pt–Pd and Pd particles with varying average properties in bimetallic catalysts.•The metal precursors influence the average properties of Pd and Pt–Pd particles.•Pt–Pd catalysts maintain high hydrogenation activity in the presence of sulfur.•Alloying degree and proportion of exposed Pt determine the catalytic performance.The tailoring of the physicochemical and catalytic properties of mono- and bimetallic Pt–Pd catalysts supported on amorphous silica–alumina was studied. Electron-energy-loss spectroscopy and extended X-ray absorption fine structure analyses indicated that bimetallic Pt–Pd and relatively large monometallic Pd particles were formed, whereas the X-ray absorption near edge structure provided direct evidence for the electronic deficiency of the Pt atoms. The heterogeneous distribution of metal particles was also shown by high-resolution transmission electron microscopy. The average structure of the bimetallic particles (Pt-rich core and Pd-rich shell) and the presence of Pd particles led to surface Pd enrichment, which was independently shown by IR spectra of adsorbed CO. The specific metal distribution, average size, and surface composition of the Pt–Pd particles depend to a large extent on the metal precursors. In the presence of NH3 ligands, Pt–Pd particles with a fairly homogeneous bulk and surface metal distribution were formed. Also, high Lewis acid site concentration of the carrier leads to more homogeneous bimetallic particles. All catalysts were active for the hydrogenation of tetralin in the absence and presence of quinoline and dibenzothiophene (DBT). Monometallic Pt catalysts had the highest hydrogenation activity in poison-free and quinoline-containing feed. When DBT was present, bimetallic Pt–Pd catalysts with the most homogenous metal distribution showed the highest activity. The higher resistance of bimetallic catalysts toward sulfur poisoning compared to their monometallic Pt counterparts results from the weakened metal–sulfur bond on the electron-deficient Pt atoms. Thus, increasing the fraction of electron-deficient Pt on the surface of the bimetallic clusters increases the efficiency of the catalyst in the presence of sulfur-containing compounds.Graphical abstractPt, Pd and Pt–Pd catalysts, supported on amorphous silica–alumina, are characterized and tested in tetralin hydrogenation. In the bimetallic catalysts, Pt–Pd and Pd particles coexist being the metal precursors decisive for the average properties of the metal particles. The bimetallic catalysts can maintain high hydrogenation activity in the presence of sulfur depending on the Pt–Pd alloying degree and proportion of exposed Pt.Download high-res image (154KB)Download full-size image
Co-reporter:Yanzhe Yu, Benjamin Fonfé, Andreas Jentys, Gary L. Haller, J.A. Rob van Veen, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (August 2012) Volume 292() pp:13-25
Publication Date(Web):1 August 2012
DOI:10.1016/j.jcat.2012.03.018
The catalytic hydrogenation of tetralin in the absence and presence of quinoline and dibenzothiophene was studied on bimetallic Pt–Pd catalysts supported on silica and amorphous silica–alumina (ASA). The proportion of Pt on the surface determined the activity given that the Pt–Pd catalyst with the highest proportion of surface Pt was the most active. In the absence of poisons, the electronegativity of the support correlated with the hydrogenation. In the presence of quinoline, the activity of the catalysts increased with the dispersion of the metal particles, whereas in the presence of dibenzothiophene, the acidity of the support controlled the activity. The observed effects of the poisons indicated the presence of two kinds of adsorption sites, that is, metal particles and Brønsted acid sites at the perimeter. The key contribution of acid sites for hydrogenation was confirmed by removing the Brønsted acid sites of the support. Reference Pt catalysts were more active than the Pt–Pd counterparts in poison-free feed. In the presence of poisons, the Pt catalysts were also more active than the bimetallic formulations with the remarkable exception of the bimetallic catalyst with the higher proportion of Pt on the surface. The bimetallic catalysts are more resistant to sulfur and nitrogen poisoning as well as to sintering. The poison resistance of bimetallic catalysts originates from the electron transference from Pt to Pd that yields weak adsorption of poisons on electron deficient Pt atoms. However, the surface coverage of Pt must be maximized to compensate the low activity of Pd-enriched catalysts.Graphical abstractThe hydrogenation of tetralin in the absence and presence of quinoline and dibenzothiophene is studied on Pt–Pd catalysts supported on amorphous silica–alumina (ASA). The proportion of Pt on the surface determines the activity. The electronegativity and acidity of the support and the average size of the metal particles are secondary influencing factors. The results point to two adsorption sites: metal and Brønsted sites at the perimeter.Download high-res image (128KB)Download full-size imageHighlights► Hydrogenation of tetralin on Pt–Pd catalysts supported on amorphous silica–alumina. ► The effects of dibenzothiophene and quinoline on the reaction are studied. ► The proportion of exposed Pt determines the performance of the catalysts. ► Acidity of the support and metal particle size are secondary influencing factors. ► The results point to two adsorption sites: metal and Brønsted sites at the perimeter.
Co-reporter:Yanzhe Yu, Benjamin Fonfé, Andreas Jentys, Gary L. Haller, J.A. Rob van Veen, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (August 2012) Volume 292() pp:1-12
Publication Date(Web):1 August 2012
DOI:10.1016/j.jcat.2012.03.017
Supported Pt–Pd catalysts were synthesized using the respective tetrammonium complexes as precursor salts. The materials used as carriers were silica and amorphous silica–alumina (ASA) with Al2O3/SiO2 wt.% ratio of 55/45, 20/80, and 5/95. The concentration of Lewis acid sites increased with the alumina content of support, whereas the concentration of Brønsted acid sites reached a maximum at 20 wt.% of alumina. According to the chemical analysis, the content of Pt and Pd was close to 0.3 and 0.5 wt.%, respectively, keeping the Pd/Pt molar ratio of 3. The average particle size slightly varied from 1.4 to 1.8 nm (TEM). The analysis of the EXAFS and CO adsorption IR characterization revealed that the general morphology of the bimetallic clusters was Pt-rich core and Pd-enriched surface. The Pt atoms in the bimetallic clusters were electron deficient due to electron transfer to Pd (XANES). Furthermore, monometallic Pd clusters coexist with those Pt–Pd particles. The proportion of surface Pt increased (from 13.9% to 16.9%) with increasing alumina content in the support. The catalyst supported on ASA(55/45) exhibited the highest coverage of Pt. The varying surface metal composition is attributed to different interaction strengths between the metal precursors and the supports during the preparation steps.Graphical abstractPt–Pd catalysts are supported on amorphous silica–alumina (ASA) of varying Al2O3/SiO2 ratio. Pd and Pt–Pd particles coexist in all the materials. Although the morphology of the bimetallic clusters corresponds to a Pt-rich core and a Pd-rich surface, the composition of the ASA carrier determines the proportion of exposed Pt.Download high-res image (128KB)Download full-size imageHighlights► Pt–Pd catalysts supported on silica and amorphous silica–alumina were characterized. ► Pt–Pd clusters form with Pt-rich core and Pd-rich surface. ► Electrons are transferred from Pt to Pd in the Pt–Pd nanoclusters. ► The alumina content of the support affects the morphology of the Pt–Pd nanoclusters.
Co-reporter:Chen Zhao, Donald M. Camaioni, Johannes A. Lercher
Journal of Catalysis (April 2012) Volume 288() pp:92-103
Publication Date(Web):1 April 2012
DOI:10.1016/j.jcat.2012.01.005
Phenol and substituted phenols are hydroalkylated and hydrodeoxygenated to bi-cycloalkanes in a tandem reaction over Pd nanoclusters supported on a large-pore molecular sieve HBEA at 473–523 K using water as solvent. The HBEA-supported Pd catalyst (metal–acid ratio: 1:22 mol/mol) optimally balances the competing rates of metal catalyzed hydrogenation as well as of solid acid-catalyzed dehydration and carbon–carbon coupling to combine hydrodeoxygenation and dimerization of phenol derivatives to C12–C18 bicycloalkanes in a single reaction sequence. A detailed kinetic study of the elementary reactions of (substituted) phenol and their potential products (cyclohexanol, cyclohexanone, and cyclohexene) demonstrates that phenol selectively reacts with the in situ generated cyclohexanol or cyclohexene on Brønsted acid sites. The acid-catalyzed alkylation of phenol with alcohol intermediates and alcohol dehydration are parallel reactions, which are subtly influenced by the competing hydrogenation reactions as well as by the presence of water as solvent. IR spectroscopy of adsorbed species and preliminary molecular modeling indicate that phenol and cyclohexanol enrichment in the large pores of zeolite HBEA is critical for the high activity and hydroalkylation selectivity.Graphical abstractPhenol and substituted phenols are hydroalkylated and hydrodeoxygenated to bi-cycloalkanes in a tandem reaction over Pd nanoclusters supported on a large-pore molecular sieve HBEA at 473–523 K using water as solvent. The HBEA-supported Pd catalyst (metal–acid ratio: 1:22 mol/mol) optimally balances the competing rates of metal catalyzed hydrogenation as well as of solid acid-catalyzed dehydration and carbon–carbon coupling to combine hydrodeoxygenation and dimerization of phenol derivatives to C12–C18 bicycloalkanes in a single reaction sequence.Download high-res image (121KB)Download full-size imageHighlights► A method has been developed on highly selective hydroalkylation and deoxygenation of substituted phenols in water. ► The acid-catalyzed alkylation and alcohol dehydration are parallel reactions in phenol hydroalkylation. ► Such competition is subtly influenced by hydrogenation and the presence of water. ► The phenol and cyclohexanol enrichment in zeolite HBEA is critical for the high activity and hydroalkylation selectivity.
Co-reporter:A. Wawrzetz, B. Peng, A. Hrabar, A. Jentys, A.A. Lemonidou, J.A. Lercher
Journal of Catalysis (5 February 2010) Volume 269(Issue 2) pp:411-420
Publication Date(Web):5 February 2010
DOI:10.1016/j.jcat.2009.11.027
Kinetically coupled reactions of glycerol in water over bifunctional Pt/Al2O3 catalysts are explored as a function of the Pt particle size and the reaction conditions. Detailed analysis of the reaction network shows that “reforming” and hydrodeoxygenation require the presence of a bifunctional catalyst, i.e., the presence of an acid–base and a metal function. The initial reaction steps are identified to be dehydrogenation and dehydration. The dehydrogenation of hydroxyl groups at primary carbon atoms is followed by decarbonylation and subsequent water gas shift or by disproportionation to the acid (and the alcohol) followed by decarboxylation. Hydrogenolysis of the C–O and C–C bonds in the alcohols does not occur under the present reaction conditions. Larger Pt particles favor hydrodeoxygenation over complete deconstruction to hydrogen and CO2.Glycerol is catalytically converted in aqueous phase over Pt/Al2O3 via bifunctional pathways involving dehydrogenation, dehydration and decarboxylation/decarbonylation. CC and CO bond hydrogenolysis does not occur.Download high-res image (37KB)Download full-size image
Co-reporter:Fabrizio Cavani, Luca Maselli, Sauro Passeri, Johannes A. Lercher
Journal of Catalysis (5 February 2010) Volume 269(Issue 2) pp:340-350
Publication Date(Web):5 February 2010
DOI:10.1016/j.jcat.2009.11.019
Reactions of phenol and methanol catalyzed by MgO have been explored by kinetic measurements and in situ IR spectroscopy combined with computational studies of sorbed molecules. On MgO, methanol partly transforms to formaldehyde above 250 °C. Adsorbed phenol forms phenolate species, and the energetically preferred mode of adsorption leads to an almost orthogonal orientation of the aromatic ring with respect to the catalyst surface. All molecules involved adsorb preferably at the corner sites of MgO (three-co-ordinated Mg atoms). The main reaction products are anisole and o-cresol, the latter dominating above 300 °C. At very low conversions, salicylic aldehyde is observed as primary reaction product, being rapidly transformed to o-cresol. It is only observed during the initial accumulation of adsorbed species on the catalyst surface, but not under steady-state conditions on a fully covered catalyst surface. Therefore, o-cresol formation starts with the reaction between phenol and formaldehyde to salicylic alcohol, which in turn is rapidly transformed to salicylic aldehyde and subsequently to o-cresol. Salicylic aldehyde may also form via the bimolecular disproportionation of salicylic alcohol to o-cresol and aldehyde. The parallel reaction to o-cresol, not involving the formation of salicylic aldehyde as intermediate, proceeds via the reduction of salicylic alcohol to o-cresol by formaldehyde. The identified mechanism may open new synthetic approaches for the production of functionalized phenol derivatives and, even more importantly, for the defunctionalization of substituted phenols potentially available at large scale from deconstructed lignin.The mechanism of the gas-phase methylation of phenol with methanol, catalyzed by basic systems, is not a conventional electrophilic ring-substitution by the activated alcohol, but instead proceeds via the intermediate formation of formaldehyde and the regio- and chemo-selective ortho-hydroxymethylation of the phenolic ring by the aldehyde.Download high-res image (116KB)Download full-size image
Co-reporter:Philipp Hauck, Andreas Jentys, Johannes A. Lercher
Catalysis Today (30 September 2007) Volume 127(Issues 1–4) pp:165-175
Publication Date(Web):30 September 2007
DOI:10.1016/j.cattod.2007.05.012
The selective catalytic reduction with aqueous solutions of urea is currently seen having the highest potential to reduce NOx and particulate emissions for commercial diesel powered vehicles. Ammonia as the actual reduction medium is formed from urea in two consecutive reactions, i.e. via the thermolysis of urea to isocyanic acid and NH3 and the catalyzed hydrolysis of HNCO over TiO2 to NH3 and CO2. A kinetic model for the hydrolysis reaction was derived for a reaction scheme comprising a set of elementary steps. To minimize the number of unknown variables in the kinetic model for the overall rate, the equilibrium constants for both reactants (HNCO and H2O) and products (NH3 and CO2) were determined from adsorption isotherms using Langmuir and multilayer adsorption models. A data set consisting of 49 data points for the rate determined at varying reactant concentrations was fitted with the kinetic model using a non-linear least mean squares regression analysis.
Co-reporter:Elvira Peringer, Chirag Tejuja, Michael Salzinger, Angeliki A. Lemonidou, Johannes A. Lercher
Applied Catalysis A: General (30 November 2008) Volume 350(Issue 2) pp:178-185
Publication Date(Web):30 November 2008
DOI:10.1016/j.apcata.2008.08.009
Co-reporter:M.F. Williams, B. Fonfé, C. Woltz, A. Jentys, J.A.R. van Veen, J.A. Lercher
Journal of Catalysis (25 October 2007) Volume 251(Issue 2) pp:497-506
Publication Date(Web):25 October 2007
DOI:10.1016/j.jcat.2007.06.010
The interactions between an oxidic support and Pt particles were explored in a series of platinum catalysts supported on amorphous silica–alumina used for the hydrogenation of tetralin and the hydrogenolysis of neopentane. The activity of the nanosized Pt particles increased with increasing intermediate electronegativity of the support for all supports with accessible Lewis acid sites. A compensation effect between the preexponential factor and the apparent energy of activation was observed for both reactions, attributed to the stronger adsorption of the reactants and intermediates on Pt induced by the increasingly electronegative supports. The product distribution in tetralin hydrogenation also points to stronger adsorption on Pt with increasing intermediate electronegativity of the support. X-ray absorption near-edge structure (XANES) findings indicate that the stronger interaction of the reactants is related to a reduced electron density on the platinum particles.
Co-reporter:M.F. Williams, B. Fonfé, C. Sievers, A. Abraham, J.A. van Bokhoven, A. Jentys, J.A.R. van Veen, J.A. Lercher
Journal of Catalysis (25 October 2007) Volume 251(Issue 2) pp:485-496
Publication Date(Web):25 October 2007
DOI:10.1016/j.jcat.2007.06.009
A series of well-defined catalysts based on platinum nanoparticles supported on amorphous silica–alumina with varying composition was prepared by incipient wetness impregnation. Quantitative structural characterization of alumina and aluminosilicate domains by 27Al (3Q) MAS NMR spectroscopy showed that the concentration of Brønsted acid sites determined by pyridine adsorption is related solely to the concentration of the aluminosilicate domain. However, only a very small fraction of the tetrahedrally coordinated aluminum led to the formation of Brønsted acid sites, because the negative charge on most of the tetrahedral Al–O was balanced by cationic aluminum oxide clusters. The preparation method led to small, uniform (0.6–0.8 nm) Pt particles on all aluminum-containing supports.
Co-reporter:Ana Hrabar, Jennifer Hein, Oliver Y. Gutiérrez, Johannes A. Lercher
Journal of Catalysis (25 July 2011) Volume 281(Issue 2) pp:325-338
Publication Date(Web):25 July 2011
DOI:10.1016/j.jcat.2011.05.017
The hydrodenitrogenation of o-propylaniline on MoS2/γ-Al2O3 and NiMoS/γ-Al2O3 catalysts proceeds via two parallel routes, i.e., direct denitrogenation (DDN) by C(sp2)–N bond cleavage to form propylbenzene and hydrogenation (HYD) of the phenyl ring to form propylcyclohexylamine, followed by C(sp3)–N bond cleavage. Coordinatively unsaturated sites (CUS) at the edges of the sulfide slabs are catalytically active for the DDN. Dibenzothiophene (DBT) decreases the DDN rate, while it is mainly converted via direct desulfurization. Adding Ni to MoS2 increases the CUS concentration and promotes the HYD but inhibits the DDN, suggesting that Ni cations are not involved in the active sites for DDN route. Catalytically active sites for the HYD route are the sites at the basal plane near the edges of the sulfide slabs (brim sites). The presence of DBT strongly increases the HYD rate on NiMoS/γ-Al2O3, increasing the electron density at the brim sites due the electron pair donor properties of DBT and biphenyl.Graphical abstractThe HDN of o-propylaniline on MoS2/γ-Al2O3 and NiMoS/γ-Al2O3 catalysts proceeds via two parallel routes on two separate sites. Direct denitrogenation (DDN) is catalyzed only by accessible Mo cations, while the sites at the basal plane near the edge of the sulfide slabs (brim sites) are active for hydrogenation (HYD).Download high-res image (57KB)Download full-size image.Highlights► Direct denitrogenation (DDN) of o-propylaniline proceeds on sulfur vacancies (CUS). ► Ni cations increase the CUS concentration but inhibits the DDN route. ► Dibenzothiophene (DBT), converted via direct desulfurization, poisons the DDN. ► DBT promotes the HYD on NiMoS, increasing the electron density at the brim sites.
Co-reporter:Chen Zhao, Jiayue He, Angeliki A. Lemonidou, Xuebing Li, Johannes A. Lercher
Journal of Catalysis (16 May 2011) Volume 280(Issue 1) pp:8-16
Publication Date(Web):16 May 2011
DOI:10.1016/j.jcat.2011.02.001
The kinetics of the catalytic hydrodeoxygenation of phenol and substituted phenols has systematically been investigated on the dual-functional catalyst system Pd/C and H3PO4 in order to better understand the elementary steps of the overall reaction. The reaction proceeds via stepwise hydrogenation of the aromatic ring, transformation of the cyclic enol to the corresponding ketone, hydrogenation of the cycloalkanone to the cycloalkanol and its subsequent dehydration as well as the hydrogenation of the formed cycloalkene. The presence of dual catalytic functions is indispensible for the overall hydrodeoxygenation. The dehydration reaction is significantly slower than the hydrogenation reaction and the keto/enol transformation, requiring a significantly larger concentration of Brønsted acid sites compared to the available metal sites for hydrogenation.Graphical abstractHydrodeoxygenation of phenol and substituted phenols over dual-functional Pd/C and H3PO4 catalysts proceeds via stepwise hydrogenation of the aromatic ring, transformation of the cyclic enol to the corresponding ketone, hydrogenation of the cycloalkanone to the cycloalkanol and its subsequent dehydration as well as the hydrogenation of the formed cycloalkene.Download high-res image (27KB)Download full-size imageResearch highlights► Dual-functional catalysts hydrodeoxygenate substituted phenols in water. ► Suitable reaction conditions lead to cycloalkane phase. ► Acid function is required to break the C–O bond via hydrolysis or dehydration.
Co-reporter:Oliver Y. Gutiérrez, Christoph Kaufmann, Ana Hrabar, Yongzhong Zhu, Johannes A. Lercher
Journal of Catalysis (13 June 2011) Volume 280(Issue 2) pp:264-273
Publication Date(Web):13 June 2011
DOI:10.1016/j.jcat.2011.03.027
The synthesis of methyl mercaptan from COS and H2 on K+ promoted MoS2 supported on silica is explored. The reaction proceeds via the disproportionation of COS to CO2 and CS2 and the consecutive hydrogenation of CS2 to CH3SH. In parallel to the disproportionation, COS also decomposes to CO and H2S. The characterization of the catalyst by means of XRD, Raman spectroscopy, and adsorption of NO suggests that two active phases, i.e., relatively pure MoS2 and K+-decorated MoS2, are present in the sulfided catalyst. The disproportionation of COS and the hydrogenation of CS2 are favored on K+-decorated MoS2; the decomposition of COS to CO is the favored route on pure MoS2. The reaction mechanisms for the decomposition of COS and the hydrogenation of CS2 are discussed.Graphical abstractThe synthesis of CH3SH on K+ promoted MoS2/SiO2 from COS and H2 proceeds through the disproportionation of COS and the consecutive hydrogenation of CS2 mainly on the K+-MoS2 phase. In parallel, COS is reduced to CO and H2S on the unpromoted MoS2 phase.Download high-res image (51KB)Download full-size imageHighlights► COS is used as educt for the synthesis of CH3SH which has received limited attention. ► COS decomposes to CO and H2S and disproportionates to CO2 and CS2, whereas CS2 hydrogenates to CH3SH. ► Two active phases are present MoS2 catalyzes COS decomposition and K+-MoS2 promotes the COS disproportionation. ► Reaction mechanisms for COS decomposition and CS2 hydrogenation are given.
Co-reporter:Virginia Roberts, Sebastian Fendt, Angeliki A. Lemonidou, Xuebing Li, Johannes A. Lercher
Applied Catalysis B: Environmental (12 March 2010) Volume 95(Issues 1–2) pp:
Publication Date(Web):12 March 2010
DOI:10.1016/j.apcatb.2009.12.010
Benzyl phenyl ether conversion in superheated water yields a broad product distribution. In addition to the hydrolysis products, phenol and benzyl alcohol, a large amount of consecutive products are formed depending on the operating conditions. The influence of Li2CO3, Na2CO3, and K2CO3 on these reactions is explored between 270 and 370 °C. It is shown that high selectivity towards hydrolysis can be achieved at low temperatures and short reaction times. At higher severities the yields of phenol and especially benzyl alcohol decrease and higher molecular weight compounds are formed by consecutive reactions. Alkali carbonates effect this distribution by decreasing the concentration of protons in the system and by providing and enhancing parallel and secondary reaction mechanisms. The yields of toluene, 2 and 4-benzyl phenol are strongly enhanced in the presence of an alkali carbonate, by formation of a cation–BPE adduct in which the ether bond is strongly polarized.
Co-reporter:Carsten Sievers, Iker Zuazo, Alexander Guzman, Roberta Olindo, Hitrisia Syska, Johannes A. Lercher
Journal of Catalysis (10 March 2007) Volume 246(Issue 2) pp:315-324
Publication Date(Web):10 March 2007
DOI:10.1016/j.jcat.2006.11.015
The formation of carbonaceous deposits and their effect on aging and deactivation of zeolite LaX during isobutane/2-butene alkylation at 348 K were investigated by stopping the reaction at different times on stream. Four stages of the reaction were identified: (1) stable alkylation, (2) deposit transformation, (3) slow deactivation, and (4) rapid deactivation. Deposits consist mostly of bicyclic compounds and branched carbenium ions, which are formed already at the beginning of the reaction and block Brønsted acid sites. During the deposit transformation, migration of smaller entities toward the pore mouth occurs. These cyclic compounds are further alkylated and lead to pore mouth plugging. In the final stage of rapid deactivation, the catalyst stops producing alkylate, and butene oligomerization is the main reaction leading to olefin desorption and massive deposit formation at the outside of the zeolite particles.
Co-reporter:Carsten Sievers, Jürgen S. Liebert, Manuel M. Stratmann, Roberta Olindo, Johannes A. Lercher
Applied Catalysis A: General (1 March 2008) Volume 336(Issues 1–2) pp:
Publication Date(Web):1 March 2008
DOI:10.1016/j.apcata.2007.09.039
Lanthanum exchanged X and Y type zeolites were prepared by ion exchange and investigated as catalysts for isobutane/2-butene alkylation. With the reactions performed in a continuously operated stirred tank reactor under industrially relevant conditions (T = 348 K, p = 20 bar, paraffin/olefin molar ratio = 10, olefin weight hourly space velocity = 0.2 h−1) the catalyst lifetime of LaX was nearly twice as long as that of LaY. Moreover, a much higher yield of octane isomers was observed with LaX. The product distributions showed that LaX had a high activity for hydride transfer and “self-alkylation” as well as a higher concentration of strong Brønsted acid sites. These differences are related to a higher residual concentration of sodium cations in LaY leading not only to less, but also weaker strong Brønsted acid sites in LaY than in LaX. The replacement of the residual sodium cations by lanthanum cations is less favorable in LaY due to the lower concentration of appropriate sites to accommodate multivalent cations.Lanthanum exchanged X and Y type zeolites were investigated as catalysts for isobutane/2-butene alkylation. The catalyst lifetime of LaX was nearly twice as long as that of LaY due to the higher strength of its Brønsted acid sites. The Brønsted acidity of LaY is markedly influenced by the presence of residual sodium cations.
Co-reporter:S. Grundner, W. Luo, M. Sanchez-Sanchez and J. A. Lercher
Chemical Communications 2016 - vol. 52(Issue 12) pp:NaN2556-2556
Publication Date(Web):2015/12/24
DOI:10.1039/C5CC08371K
Cu-Exchanged zeolites are known as active materials for methane oxidation to methanol. However, understanding of the formation of Cu active species during synthesis, dehydration and activation is fragmented and rudimentary. We show here how a synthesis protocol guided by insight in the ion exchange elementary steps leads to highly uniform Cu species in mordenite (MOR).
Co-reporter:Oliver Y. Gutiérrez, Yanzhe Yu, Robin Kolvenbach, Gary L. Haller and Johannes A. Lercher
Catalysis Science & Technology (2011-Present) 2013 - vol. 3(Issue 9) pp:NaN2372-2372
Publication Date(Web):2013/07/02
DOI:10.1039/C3CY00189J
Pt supported on sulfated zirconia (SZ) and amorphous silica alumina (ASA) was explored for the hydrogenation of tetralin. Pt/ASA had a higher concentration and strength of acid sites than the SZ-supported counterpart, for which the acid site concentration of the carrier decreases after catalyst synthesis due to sulfur elimination. The strong acidity of Pt/ASA caused higher deactivation due to coke deposition, while Pt/SZ was comparatively stable. Pt/ASA also exhibited higher porosity and Pt dispersion than Pt/SZ. The adsorption of CO, the XANES analysis and gravimetric sorption of benzene showed that the Pt particles on SZ were strongly electron deficient, a feature that has been speculated to be associated with the electron withdrawing effect of sulfate groups. For tetralin hydrogenation, Pt/SZ was more active than Pt/ASA above 423 K while the opposite was observed at 413 K. The apparent activation energies were 98 and 45 kJ mol−1 on Pt/SZ and Pt/ASA, respectively. Pt/ASA was more active than Pt/SZ in the presence of quinoline, while Pt/SZ retains the highest activity in the presence of dibenzothiophene (with or without quinoline). The lower apparent activation energy on Pt/ASA in the absence of quinoline or dibenzothiophene and its higher activity in the presence of quinoline were attributed to its strong Brønsted acidity, promoting the hydrogenation at the perimeter of Pt particles. The higher sulfur resistance of Pt/SZ was attributed to the electron deficiency of the supported Pt particles. In this respect, surface sulfate anions induce stronger electron deficiency in supported Pt than the acidity of the support.
Co-reporter:Jiechen Kong, Mingyuan He, Johannes A. Lercher and Chen Zhao
Chemical Communications 2015 - vol. 51(Issue 99) pp:NaN17583-17583
Publication Date(Web):2015/09/28
DOI:10.1039/C5CC06828B
The utilization of lignin as a fuel precursor has attracted attention, and a novel and facile process has been developed for one-pot conversion of lignin into cycloalkanes and alkanes with Ni catalysts under moderate conditions. This cascade hydrodeoxygenation approach may open the route to a new promising technique for direct liquefaction of lignin to hydrocarbons.
Co-reporter:Chen Zhao, Yuan Kou, Angeliki A. Lemonidou, Xuebing Li and Johannes A. Lercher
Chemical Communications 2010 - vol. 46(Issue 3) pp:NaN414-414
Publication Date(Web):2009/11/18
DOI:10.1039/B916822B
A simple, green, cost- and energy-efficient route for converting phenolic components in bio-oil to hydrocarbons and methanol has been developed, with nearly 100% yields. In the heterogeneous catalysts, RANEY® Ni acts as the hydrogenation catalyst and Nafion/SiO2 acts as the Brønsted solid acid for hydrolysis and dehydration.
Co-reporter:Sonja A. Wyrzgol, Susanne Schäfer, Sungsik Lee, Byeongdu Lee, Marcel Di Vece, Xuebing Li, Sönke Seifert, Randall E. Winans, Martin Stutzmann, Johannes A. Lercher and Stefan Vajda
Physical Chemistry Chemical Physics 2010 - vol. 12(Issue 21) pp:NaN5595-5595
Publication Date(Web):2010/04/27
DOI:10.1039/B926493K
The preparation, characterization and catalytic reactivity of a GaN supported Pt catalyst in the hydrogenation of ethene are presented in this feature article, highlighting the use of in situ characterization of the material properties during sample handling and catalysis by combining temperature programmed reaction with in situ grazing incidence small-angle X-ray scattering and X-ray absorption spectroscopy. The catalysts are found to be sintering resistant at elevated temperatures as well as during reduction and hydrogenation reactions. In contrast to Pt particles of approximately 7 nm diameter, smaller particles of 1.8 nm in size are found to dynamically adapt their shape and oxidation state to the changes in the reaction environment. These smaller Pt particles also showed an initial deactivation in ethene hydrogenation, which is paralleled by the change in the particle shape. The subtle temperature-dependent X-ray absorbance of the 1.8 nm sized Pt particles indicates that subtle variations in the electronic structure induced by the state of reduction by electron tunnelling over the Schottky barrier between the Pt particles and the GaN support can be monitored.
Co-reporter:John H. Ahn, Robin Kolvenbach, Sulaiman S. Al-Khattaf, Andreas Jentys and Johannes A. Lercher
Chemical Communications 2013 - vol. 49(Issue 90) pp:NaN10586-10586
Publication Date(Web):2013/09/23
DOI:10.1039/C3CC46197A
An increase in p-xylene selectivity was observed without losing the catalytic activity over novel mesoporous nano-sized ZSM5 crystals covered with an external SiO2 overlayer created by deposition of tetraethyl orthosilicate.
Molybdenum nickel tungsten sulfide
Benzenamine, N,N-di-2-cyclohexen-1-yl-
Benzenamine, N-(2-cyclohexen-1-ylmethylene)-4-(1-methylethyl)-
Benzenamine, 4-(1-methylethyl)-N-(1-phenylethylidene)-
Benzenamine, 4-fluoro-N-(1-phenylethylidene)-
Piperidine, 2-methylene-
Benzenamine, N-2-cyclohexen-1-yl-3-methoxy-
Benzoic acid, 4-(2-cyclohexen-1-ylamino)-, ethyl ester
Benzenamine, N-2-cyclohexen-1-yl-4-(trifluoromethyl)-
Benzenamine, N-2-cyclohexen-1-yl-2-methoxy-